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 46 毫秒
1.
The objective of this study was to investigate the pH-dependent solubility and dissolution of weakly basic Biopharmaceutical Classification Systems (BCS) class II drugs, characterized by low solubility and high permeability, using carvedilol, a weak base with a pK a value of 7.8, as a model drug. A series of solubility and in vitro dissolution studies was carried out using media that simulate the gastric and intestinal fluids and cover the physiological pH range of the GI from 1.2 to 7.8. The effect of ionic strength, buffer capacity, and buffer species of the dissolution media on the solubility and dissolution behavior of carvedilol was also investigated. The study revealed that carvedilol exhibited a typical weak base pH-dependent solubility profile with a high solubility at low pH (545.1–2591.4 μg/mL within the pH range 1.2–5.0) and low solubility at high pH (5.8–51.9 μg/mL within the pH range 6.5–7.8). The dissolution behavior of carvedilol was consistent with the solubility results, where carvedilol release was complete (95.8–98.2% released within 60 min) in media simulating the gastric fluid (pH 1.2–5.0) and relatively low (15.9–86.2% released within 240 min) in media simulating the intestinal fluid (pH 6.5–7.8). It was found that the buffer species of the dissolution media may influence the solubility and consequently the percentage of carvedilol released by forming carvedilol salts of varying solubilities. Carvedilol solubility and dissolution decreased with increasing ionic strength, while lowering the buffer capacity resulted in a decrease in carvedilol solubility and dissolution rate.  相似文献   

2.
The lytic effect of lysozyme on Streptococcus faecalis ATCC 9790 was studied by spectrophotometry and electron microscopy and it was found to be highly dependent on the ionic strength of the suspending media and on the ratio lysozyme to bacterial cell mass. When 7.2 X 10(8) bacteria/mL are exposed to 0.4 mg/mL of lysozyme in media with low ionic strength, the enzyme is bound in great amounts, as deduced from protein determinations and sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS--PAGE); the binding prevents bacteriolysis in spite of the removal of the cell wall. Extensive lysis of S. faecalis could be obtained by reducing the ratio of lysozyme to bacterial cell mass. Stabilization of S. faecalis by lysozyme was also observed when exponential phase cells incubated under conditions that promote spontaneous autolysis (incubation in 0.05 M tris(hydroxymethyl)aminomethane buffer, pH 8.0, ionic strength = 0.01675) do not lyse and do not leak material which absorbs at 260 nm when lysozyme was present at the highest concentration.  相似文献   

3.
A computer program that allows the preparation of buffers containing known concentrations of metal-ligand complexes at defined pH values and temperatures is described. Ligands are defined as compounds that bind metals and may include AMP, ADP, ATP, GMP, GDP, GTP, EGTA, EDTA, BAPTA, phosphate, sulfate, chloride, monocarboxylic acids, dicarboxylic acids, organophosphates, and/or citric acid. Metals may include sodium, potassium, magnesium, calcium, and/or manganese. The program uses association constants corrected for temperature and ionic strength so that solutions between 0 and 40 degrees C and between pH values of 4 and 10 can be defined. The program can perform the following: (i) calculate the concentration of all metal-ligand complexes when total metal and total ligand concentrations are known, (ii) calculate the concentration of metal ion required to make a solution of known free metal ion concentration when total ligand concentrations are known, (iii) calculate the concentration of ligand required to make a solution of known free metal ion concentration when total metal concentrations are known, and (iv) calculate the total concentrations of metal and ligand required to make a buffer of known metal-ligand concentration. Options i-iii are useful for making buffers of defined free metal ion concentrations; option iv is useful for making buffers of defined metal-nucleotide concentrations.  相似文献   

4.
The stability of two nitrosoureas, tauromustine and lomustine, has been investigated in different media and buffers. All media tested, except Leibovitz's L-15 medium, significantly increased the degradation rate of the investigated nitrosoureas at pH 7.4. Sodium bicarbonate seems to be the cause of the observed increase of the degradation rate, since it provides the main buffering capacity of all the media except for Leibovitz's L-15 medium, which is based on phosphate buffer. Other ingredients in the media, such as amino acids, vitamins, and inorganic salts, or the ionic strength of a buffer, did not have any major effect on the degradation rate of the nitrosoureas. These results suggest that media containing carbonated buffer should be avoided when the anti-tumor effect of nitrosoureas is to be investigated in different cell cultures.  相似文献   

5.
6.
Formation of complexes between bovine beta-lactoglobulins (BLG) and long-chain fatty acids (FAs), effect of complex formation on protein stability, and effects of pH and ionic strength on both complex formation and protein stability were investigated as a function of pH and ionic strength by electrophoretic techniques and NMR spectroscopy. The stability of BLG against unfolding is sharply affected by the pH of the medium: both A and B BLG variants are maximally stabilized against urea denaturation at acidic pH and against SDS denaturation at alkaline pH. The complexes of BLGB with oleic (OA) and palmitic acid (PA) appear more stable than the apoprotein at neutral pH whereas no differential behavior is observed in acidic and alkaline media. PA forms with BLG more stable complexes than OA. The difference between the denaturant concentration able to bring about protein unfolding in the holo versus the apo forms is larger for urea than for SDS treatment. This evidence disfavors the hypothesis of strong hydrophobic interactions being involved in complex formation. Conversely, a significant contribution to FA binding by ionic interactions is demonstrated by the effect of pH and of chloride ion concentration on the stoichiometry of FA.BLG complexes. At neutral pH in a low ionic strength buffer, one molecule of FA is bound per BLG monomer; this ratio decreases to ca. 0.5 per monomer in the presence of 200 mM NaCl. The polar heads of bound FA appear to be solvent accessible, and carboxyl resonances exhibit an NMR titration curve with an apparent pK(a) of 4.7(1).  相似文献   

7.
1. The effect of pH change on the reconstitution of aspartate aminotransferase (EC 2.6.1.1), i.e. the reactivation of the apoenzyme with coenzyme (pyridoxal phosphate and pyridoxamine phosphate), was studied in the pH range 4.2-8.9 by using three buffer systems at concentrations ranging from 0.025 to 0.1m. 2. Although the profile of the reconstitution rate-pH curve in the range pH5.2-6.8 (covered by sodium cacodylate-HCl buffer) reflects the influence of the H(+) concentration on the reconstitution process, the profile of the curve in the pH ranges 4.2-5.6 and 7.2-8.25 (covered respectively by sodium acetate-acetic acid and Tris-HCl buffers) appears to be influenced by the ionic strength of the buffer. 3. The reconstitution is also influenced by univalent inorganic ions such as halide ions and, to a lesser extent, alkali metal ions, which are known to alter the water structure.  相似文献   

8.
An algorithm is presented for the calculation of metal ion concentrations from given total metal concentrations (and vice versa) in physiological media containing metal-chelating compounds. In such media, conditions differ from those used for stability constant determination of metal-chelator equilibria; therefore calculated metal ion concentrations are incorrect. We recompute stability constants to reflect the effects of ionic strength and temperature of physiological solutions. Twelve different equilibria can be considered per metal-chelator pair. The computer program also calculates the contribution of ionized species of metals, chelator, complexes and pH buffers to ionic strength. Measurements with a Ca-selective electrode and with fura-2 show that calculated ionic Ca2+ concentrations are correct from 10 nM up to the millimolar range. The importance of the correct calculation of metal ion concentrations in physiological experiments is demonstrated by data, and derived kinetic parameters, on Na+/Ca2+ exchange and the ATP-dependent Ca2+ pump of enterocyte plasma membrane vesicles. The program is written in Turbo Pascal and will run on IBM-compatible computers. It is menu-driven and supports the use of a Microsoft mouse.  相似文献   

9.
Precipitation of collagens by polyethylene glycols   总被引:2,自引:0,他引:2  
Types I, II, and III collagens are readily precipitated at neutral pH by polyethylene glycols (PEG). As the molecular weight fraction of the polyethylene glycols increases, they become more effective as precipitants on a weight basis. The amount of PEG required for precipitation depends on the pH, the ionic strength, and the nature of the buffer or salts present. In tissue culture media, low concentrations of collagens and procollagens may be quantitatively precipitated and readily collected by low-speed centrifugation. Polyethylene glycol precipitation can be used to obtain collagens and procollagens from tissue culture media at either analytical or preparative scale, and since the polyethylene glycols do not bind to collagens, the precipitates may be further analyzed directly by chromatographic or electrophoretic methods.  相似文献   

10.
Diafiltration is used for final formulation of essentially all biotherapeutics. Several studies have demonstrated that buffer/excipient concentrations in the final diafiltered product can be different than that in the diafiltration buffer due to interactions between buffer species and the protein product. However, recent work in our lab has shown variations in solution pH that are largely independent of the protein concentration during the first few diavolumes. Our hypothesis is that these pH variations are due to nonidealities in the acid‐base equilibrium coefficient. A model was developed for the diafiltration process accounting for the ionic strength dependence of the pKa. Experimental results obtained using phosphate and histidine buffers were in excellent agreement with model predictions. A decrease in ionic strength leads to an increase in the pKa for the phosphate buffer, causing a shift in the solution pH, even under conditions where the initial feed and the diafiltration buffer are at the same pH. This effect could be eliminated by matching the ionic strength of the feed and diafiltration buffer. The experimental data and model provide new insights into the factors controlling the pH profile during diafiltration processes. © 2017 American Institute of Chemical Engineers Biotechnol. Prog., 33:1555–1560, 2017  相似文献   

11.
Osmotic pressure studies were carried on tobacco mosaic virus (TMV) and its components, protein and RNA, as well as on bis(3,3′-aminopropyl)amine, reported to be present in TMV preparations. Solvents were phosphate and barbital buffers at different values of pH and ionic strength. Measurements were made at room temperature. The Donnan effect was exhibited by TMV protein in phosphate buffer of 0.01 ionic strength at pH values ranging between 5.8 and 7.5. The observed values of the Donnan effect at pH 5.8 and 5.97 were in reasonable agreement with theoretical values calculated from the charge obtained by hydrogen ion titration. TMV-RNA in phosphate buffer at pH 7.5 and ionic strength 0.01 did not exhibit more than 1% of the expected Donnan effect. This is explained tentatively as the result of firm binding of metal ions. Negative values of osmotic pressure were observed with bis(3,3′-aminopropyl)amine. Similar anomalous osmosis was sometimes observed with TMV protein and with TMV. In agreement with earlier observations, TMV did not exhibit the Donnan effect in phosphate buffer of 0.01 ionic strength at pH values ranging from 5.5 to 8.0. However, TMV dialysed extensively in the presence of EDTA at pH 8.5 and TMV produced by reconstitution of purified protein and RNA did exhibit the Donnan effect in both phosphate and barbital buffers. The magnitude was of the same order as that calculated from the net charge determined by hydrogen ion titration. When reconstituted TMV, which did exhibit Donnan effect, was treated with calcium ions, the effect was abolished.  相似文献   

12.
This paper describes a stack for the Apple MacintoshlTM, HypercardTMenvironment that facilitates the calculation of a set of thermodynamically-correctedpH buffers. Presented in the familiar Macintosh mouse-basedenvironment, the program allows comprehensive buffer designand gives the user full control of buffer species, temperature,ionic strength, pH and choice of counter-ion to maintain theionic strength at a fixed value. Addition of new buffers tothe stack is straightforward and the recipes that are generated,including the weights of the various components, can be displayedon the screen or saved to a disk file for subsequent printing. Received on May 18, 1988; accepted on July 15, 1988  相似文献   

13.
After dialysis against 10 mM-Tris-acetate (pH 8.5), vimentin that has been purified in the presence of urea is present in the form of tetrameric 2 to 3 nm X 48 nm rods known as protofilaments. These building blocks in turn polymerize into intermediate filaments (10 to 12 nm diameter) when they are dialyzed against a solution of physiological ionic strength and pH. By varying the ionic conditions under which polymerization takes place, we have identified two classes of assembly intermediates whose structures provide clues as to how an intermediate filament may be constructed. The structure of the first class, seen when assembly takes place at 10 to 20 mM-salt at pH 8.5, strongly suggests that one of the initial steps of filament assembly is the association of protofilaments into pairs with a half-unit axial stagger. Increasing the ionic strength of the assembly buffer leads to the emergence of short, full-width intermediate filaments at approximately 50 mM-salt at pH 8.5. In the presence of additional protofilaments, these short filaments elongate to many micrometers when the ionic strength and pH are further adjusted to physiological levels. The electron microscope images of the assembly intermediates suggest that vimentin-containing intermediate filaments are made up of eight protofilaments, assembled such that there is an approximately 22 nm axial stagger between neighboring protofilaments. We propose that this half-unit staggering of protofilaments is a fundamental feature of intermediate filament structure and assembly, and that it could account for the 20 to 22 nm axial repeat seen in all intermediate filaments examined so far.  相似文献   

14.
Buffers of different pH and ionic strength were employed as coating buffers for antigen adsorption to microtitre plates. Their efficiency for coating plates with rinderpest virus (RPV) and foot-and-mouth disease virus (FMDV) antigens was studied by ELISA with polyclonal and monoclonal antibody preparations. While the adsorption and detection of RPV antigen with polyclonal antiserum was highly dependent on the ionic strength and pH of coating buffer, adsorption of antigenically active FMDV antigen was relatively unaffected by the buffering conditions. Both antigens were adsorbed optimally in 0.01 M phosphate buffer, pH 8.0. When monoclonal antibodies were used to detect antigen, there was a greater degree of dependence on the coating buffer than that found with polyclonal antisera. Moreover, when they were used to detect antigen adsorbed under several buffering conditions, monoclonal antibodies showed a variety of preferred buffers. The usefulness of this differential reactivity in distinguishing epitope specificity is demonstrated.  相似文献   

15.
Two methods of isolating Fab- and Fc-fragments from mouse immunoglobulin G1 are presented. The first method involves fractionation of papain protein hydrolysate on a column with DEAE- (or DE-32)-cellulose adjusted with 0.005 M K-phosphate buffer, pH 8. The Fab-fragment was eluted from the column with the starting buffer. The Fc-fragment was eluted, with the buffer ionic strength being increased to 0.4 M. Another method involves protein fractionation on an ion exchanger adjusted with 0.004 M Tris-H3PO4 buffer, pH 8.5. All the protein was column bound. The Fab-fragment was eluted with 0.04 M Tris-buffer containing a 0.004 M mixture of K-phosphates, pH 8.6. The Fc-fragment was eluted, with ionic strength being raised to 0.4 M with phosphates. As none of the methods assures isolation of absolutely pure Fab- or Fc-fragments, it is requird that cross absorption of antisera with respective immunosorbents may be carried on in order to obtain monospecific antisera to these fragments.  相似文献   

16.
R O Ryan  L D Hicks  C M Kay 《FEBS letters》1990,267(2):305-310
Hydrodynamic studies conducted in the analytical ultracentrifuge provided evidence for two populations of lipid transfer particle (LTP) when centrifuged in a buffer solution containing 10 mM Tris, pH 8.0/100 mM KCl. The apparent sedimentation coefficients of the two species was 23.3 S and 15.3 S. Upon changing the buffer pH to 7.0 or 5.7, two species of LTP were still present but the ratio of their relative abundance was altered. When the KCl concentration in the buffer was lowered to 50 mM the sample sedimented as a single species with an apparent S20,w of 22.9 S. In higher ionic strength buffers (10 mM succinate, pH 5.7/500 mM KCl) LTP sedimented with an apparent S20,w of 14.8 S. Further experiments revealed that these two forms are interconvertable as a function of buffer ionic strength. Given previous estimates of the molecular size of LTP we concluded that the slower sedimenting peak observed at high ionic strength represents monomeric LTP while the faster sedimenting material observed at low ionic strength is likely to be an aggregated state of LTP. This interpretation is supported by molecular weight determinations made by sedimentation equilibrium experiments conducted in 10 mM succinate, pH 5.7/500 mM KCl which yielded a particle Mr = 887,000. Circular dichroism spectra of monomeric LTP sample revealed 6% alpha-helix, 49% beta-sheet, 7% beta-turn and 35% random coil while aggregated LTP contained 13% alpha-helix, 66% beta-sheet and 21% random coil. The transfer activity of the two LTP forms was assayed and found to be the same indicating that either the state of LTP aggregation did not affect transfer activity or that upon exposure to a large excess of lipoprotein substrate disaggregation, without loss of activity, occurs.  相似文献   

17.
The mobilities of coenzyme A and coenzyme A derivatives of cholate, chenodeoxycholate, deoxycholate, lithocholate, and their 5 alpha analogs were studied in reversed-phase high-performance liquid chromatography. With a C18 Radial-PAK A cartridge (10-micron particles) and a solvent mixture of 2-propanol/10 mM phosphate buffer (pH 7.0, 140:360), separation of the chenodeoxycholyl and deoxycholyl coenzyme A derivatives was not observed. An increase in ionic strength of the buffer to 50 mM afforded separation, which was markedly augmented with a C18 Radial-PAK A cartridge with 5-micron particles. Lowering the pH of the buffer to 5.5 did not materially change the separations regardless of the ionic strength. Quantitation was carried out to a lower level of 8.5 X 10(-12) mol.  相似文献   

18.
The HPLC separation of the R,S and S,R enantiomers of pyrrolidinyl norephedrine on immobilized alpha-1 glycoprotein (AGP) was investigated. Conditions for the separation were varied using a premixed mobile phase containing an ammonium phosphate buffer and an organic modifier. The influence of mobile phase pH, ionic strength, organic modifier composition, modifier type, and temperature on the chiral selectivity and retention were investigated. The presented data demonstrate that independent phenomena govern the enantioselectivity and retention. Retention is a function of both ion exchange equilibria and hydrophobic adsorption. Thermodynamic data derived from van't Hoff plots illustrates that while enantioselectivity is also enthalpically driven, the magnitude of the enthalpy term is governed by pH. Enantioselectivity has little dependence on ionic strength. Hydrophobic interactions appear to foster hydrogen bonding interactions; the two appear to be mutually responsible for chiral selectivity. The chiral selectivity decreases as the pH is decreased and increases with mobile phase buffer strength.  相似文献   

19.
The experimental validation refers to the computer program reported in the companion paper, able to simulate the course of pH, buffering power (beta) and ionic strength (I) of polyprotic buffers (either singly or in a mixture) titrated over any pH range. With simple oligoamines (up to five nitrogens) it is shown that it is impossible to generate linear pH gradients in the pH 4-10 interval, unless they are mixed in appropriate ratios. With pentaethylene hexamine, when used alone, it is possible to create a linear pH 4-10 interval, provided the molarity ratios are altered in the two chambers of the gradient mixer. The general rule operating for generation of linear pH intervals is constancy of buffering power throughout the titration. Local minima of beta produce steeper gradients, while local beta maxima flatten it. The ideal delta pK to arrange for linear pH gradients during titration is centred around 1 pH unit; thus polyprotic buffers with very large delta pK values (e.g., EDTA) appear to be totally useless for this purpose. The present computing algorithms should be quite efficient for optimizing existing buffer recipes for chromatofocusing or ampholyte displacement chromatography or for creating new, properly tailored, buffer mixtures.  相似文献   

20.
We present plasmonic optical trapping of micron-sized particles in biologically relevant buffer media with varying ionic strength. The media consist of 3 cell-growth solutions and 2 buffers and are specifically chosen due to their widespread use and applicability to breast-cancer and angiogenesis studies. High-precision rheological measurements on the buffer media reveal that, in all cases excluding the 8.0 pH Stain medium, the fluids exhibit Newtonian behavior, thereby enabling straightforward measurements of optical trap stiffness from power-spectral particle displacement data. Using stiffness as a trapping performance metric, we find that for all media under consideration the plasmonic nanotweezers generate optical forces 3–4x a conventional optical trap. Further, plasmonic trap stiffness values are comparable to those of an identical water-only system, indicating that the performance of a plasmonic nanotweezer is not degraded by the biological media. These results pave the way for future biological applications utilizing plasmonic optical traps.  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