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.
Arginine deiminase (ADI, EC 3.5.3.6) is a potential antitumor drug for the treatment of arginine‐auxotrophic tumors such as hepatocellular carcinomas (HCCs) and melanomas, and studies on human lymphatic leukemia cell lines have confirmed that ADI has antiangiogenic activity. Recent studies showed that a combination of taxane and ADI‐PEG20, which induces caspase‐independent apoptosis, is more effective than taxane monotherapy for prostate cancer. The main limitation of ADI from Pseudomonas plecoglossicida (PpADI) and of many other ADI enzymes lies in their pH‐dependent activity profile. PpADI has a pH optimum at 6.5 and a pH shift from 6.5 to 7.5 results in an ~80 % activity drop (the pH of human plasma is 7.35 to 7.45). In 2010, we reported a proof of concept for ADI engineering by directed evolution that resulted in variant M2 (K5T/D44E/H404R). M2 has a pH optimum of pH 7.0, a fourfold higher kcat value than the wild‐type PpADI (pH 7.4, 0.5 M phosphate buffer), and an increased Km value for substrate arginine. In our latest work, variants M5 (K5T/D38H/D44E/A128T/H404R) and M6 (K5T/D38H/D44E/A128T/E296K/H404R) were generated by directed evolution by employing PBS buffer (pH 7.4), which mimics physiological conditions. The S0.5 value of parent M3 (K5T/D44E/A128T/H404R) decreased from 2.01 to 1.48 mM (M5) and 0.81 mM (M6). The S0.5 value of M6 (0.81 mM ) is lower than that of wild‐type PpADI (1.30 mM ); the kcat values improved from 0.18 s?1 (wild‐type PpADI) to 17.56 s?1 (M5, 97.6‐fold) and 11.64 s?1 (M6, 64.7‐fold).  相似文献   

2.
Gene‐inactivation experiments have indicated that the putative prenyltransferase XptB from Aspergillus nidulans was likely to be responsible for the prenylation of 1,7‐dihydroxy‐6‐methyl‐8‐hydroxymethylxanthone. Recently, it was suggested that this enzyme might also accept as substrate the benzophenone arugosin H, which is assumed to be a precursor of prenylated xanthones. In this study, five benzophenones and ten xanthones were incubated with purified recombinant XptB in the presence of dimethylallyl diphosphate (DMAPP). XptB accepted four xanthones as substrates, including the proposed natural substrate, and catalysed regiospecific O‐prenylations at C‐7 of the xanthone core. Km values in the range of 0.081–1.1 mM and turnover numbers (kcat) between 0.02 and 0.5 s?1 were determined for the accepted xanthones. The kinetic parameters for DMAPP were found to be 0.024 mM (Km) and 0.13 s?1 (kcat). Arugosin H was not accepted by XptB under the tested conditions. XptB was relatively specific towards its prenyl donor and did not accept geranyl or farnesyl diphosphate as substrate. Mn2+ and Co2+ strongly enhanced XptB activity (up to eightfold); this has not been reported before for prenyltransferases of the DMATS superfamily.  相似文献   

3.
Alkaline esterase (carboxylic‐ester hydrolases; EC 3.1.1.1) extracted from germinated soybean seeds (Glycine max) was purified approximately 3.6 times by chromatography in a DEAE‐cellulose anion exchange column and filtration in Sephadex G100 gel. The molecular mass of the enzyme was estimated at 45 kDa by gel electrophoresis (SDS‐PAGE). The purified enzyme showed a specific activity of 5.6 U mg?1 using p‐nitrophenyl butyrate as substrate. The esterase showed optimal activity at 47 °C in moderately alkaline pH, low stability in temperatures higher than 50 °C, and high stability at pH values between 6 and 9.5. The Ca2+ and Co2+ ions proved to have a positive effect on enzyme activity; however, Hg2+ completely inhibited esterase activity. Using p‐nitrophenyl butyrate as substrate, the enzyme showed a Km of 0.39 mM, Vmax of 31.5 mM mg?1 min?1 and kcat 7.60 × 106 s?1. Regarding substrate affinity, the enzyme showed greater activity for substrates containing short‐chain fatty acids, especially p‐nitrophenyl acetate. Such characteristics give the enzyme great potential for application in the production of low molecular weight esters, in the food industry, and in chemical products. This enzyme is another new member of the family of lipases and esterases from vegetable seeds with high activity and stability in alkaline pH.  相似文献   

4.
A gene from the marine bacterium Stenotrophomonas maltophilia encodes a 38.6 kDa FAD‐containing flavoprotein (Uniprot B2FLR2) named S. maltophilia flavin‐containing monooxygenase (SMFMO), which catalyses the oxidation of thioethers and also the regioselective Baeyer–Villiger oxidation of the model substrate bicyclo[3.2.0]hept‐2‐en‐6‐one. The enzyme was unusual in its ability to employ either NADH or NADPH as nicotinamide cofactor. The KM and kcat values for NADH were 23.7±9.1 μM and 0.029 s?1 and 27.3±5.3 μM and 0.022 s?1 for NADPH. However, kcat/KM value for the ketone substrate in the presence of 100 μM cofactor was 17 times greater for NADH than for NADPH. SMFMO catalysed the quantitative conversion of 5 mM ketone in the presence of substoichiometric concentrations of NADH with the formate dehydrogenase cofactor recycling system, to give the 2‐oxa and 3‐oxa lactone products of Baeyer–Villiger reaction in a ratio of 5:1, albeit with poor enantioselectivity. The conversion with NADPH was 15 %. SMFMO also catalysed the NADH‐dependent transformation of prochiral aromatic thioethers, giving in the best case, 80 % ee for the transformation of p‐chlorophenyl methyl sulfide to its R enantiomer. The structure of SMFMO reveals that the relaxation in cofactor specificity appears to be accomplished by the substitution of an arginine residue, responsible for recognition of the 2′‐phosphate on the NADPH ribose in related NADPH‐dependent FMOs, with a glutamine residue in SMFMO. SMFMO is thus representative of a separate class of single‐component, flavoprotein monooxygenases that catalyse NADH‐dependent oxidations from which possible sequences and strategies for developing NADH‐dependent biocatalysts for asymmetric oxygenation reactions might be identified.  相似文献   

5.
The inverse microemulsion copolymerization of acrylamide and butyl acrylate initiated by ammonium peroxodisulphate, a water‐soluble initiator, and stabilized by anionic emulsifiers sodium bis‐2‐ethylhexylsulfosuccinate and sodium dodecylsulphate (SDS) has been studied. An increase of SDS concentration was observed to increase both the rate of polymerization and the particle size. The average number of radicals per particle (n?) is much below 0.5, which indicates desorption of monomeric radicals from polymer particles. The exit (desorption) rate constants kdes (cm2 s?1) and kdes (s?1) were evaluated as a function of SDS concentration (or the particle size) according to the Ugelstad/O'Toole (I), Nomura (II) and Gilbert (III) models. The value of kdes (s?1) decreases with increasing particle size (or SDS concentration) for all three (I, II and III) models. A complex trend appears for kdes (cm2 s?1): the Ugelstad/O'Toole model estimates a decrease, the Nomura model finds no variation and the Gilbert model estimates even a slight increase in kdes with increasing SDS concentration. Copyright © 2006 Society of Chemical Industry  相似文献   

6.
The substrate scope of the flavoprotein alditol oxidase (AldO) from Streptomyces coelicolor A3(2), recombinantly produced in Escherichia coli, was explored. While it has been established that AldO efficiently oxidizes alditols to D ‐aldoses, this study revealed that the enzyme is also active with a broad range of aliphatic and aromatic alcohols. Alcohols containing hydroxy groups at the C‐1 and C‐2 positions like 1,2,4‐butanetriol (Km=170 mM, kcat=4.4 s−1), 1,2‐pentanediol (Km=52 mM, kcat=0.85 s−1) and 1,2‐hexanediol (Km=97 mM, kcat=2.0 s−1) were readily accepted by AldO. Furthermore, the enzyme was highly enantioselective for the oxidation of 1,2‐diols [e.g., for 1‐phenyl‐1,2‐ethanediol the (R)‐enantiomer was preferred with an E‐value of 74]. For several diols the oxidation products were determined by GC‐MS and NMR. Interestingly, for all tested 1,2‐diols the products were found to be the α‐hydroxy acids instead of the expected α‐hydroxy aldehydes. Incubation of (R)‐1‐phenyl‐1,2‐ethanediol with 18O‐labelled water (H218O) revealed that a second enzymatic oxidation step occurs via the hydrate product intermediate. The relaxed substrate specificity, excellent enantioselectivity, and independence of coenzymes make AldO an attractive enzyme for the preparation of optically pure 1,2‐diols and α‐hydroxy acids.  相似文献   

7.
Enzyme efficiency results from the cooperation of functional groups in the catalytic site. In order to mimic a natural enzyme, a definite 3D scaffold must be carefully designed so that the functional groups can work cooperatively. During the HIV‐1 fusion process, the gp41 N‐ and C‐terminal heptad repeat regions form a coiled‐coil six‐helical bundle (6HB) that brings the viral and target cell membranes into close proximity for fusion. We used 6HB as the molecular model for a novel scaffold for the design of an artificial enzyme, in which the modified C34 and N36 peptides formed a unique 6HB structure through specific molecular recognition, and the position and orientation of the side‐chain groups on this scaffold were predictable. The histidine modified 6HB C34H13/20/N36H15/22 showed enzyme‐like hydrolytic activity towards p‐nitrophenyl acetate (PNPA; kcat/KM=3.66 M ?1 s?1) through the cooperation of several inter‐ or intrahelical imidazole groups. Since the catalytic activity of 6HB depends on the C‐ and N‐peptide assembly, either HIV fusion inhibitors that can compete with the formation of catalytic 6HB or denaturants that can destroy the ordered structure were able to modulate its activity. Further engineering of the solvent‐exposing face with Glu?‐Lys+ salt bridges enhanced the helicity and the stability of 6HB. As a result, the population and stability of cooperative catalytic units increased. In addition, the Glu?‐Lys+‐stabilized 6HB SC35H13/20/N36H15/22 had increased catalytic efficiency (kcat/KM=6.30 M ?1 s?1). A unique 6HB system was specifically assembled and provided a scaffold sufficiently stable to mimic the function of enzymes or other biomolecules.  相似文献   

8.
Natural and synthetic unsaturated glucuronides were tested as substrates for Clostridium perfringens unsaturated glucuronyl hydrolase to probe its mechanism and to guide inhibitor design. Of the natural substrates, a chondroitin disaccharide substrate with sulfation of the primary alcohol on carbon 6 of its N‐acetylgalactosamine moiety was found to have the highest turnover number of any substrate reported for an unsaturated glucuronyl hydrolase, with kcat=112 s?1. Synthetic aryl glycoside substrates with electron‐withdrawing aglycone substituents were cleaved more slowly than those with electron‐donating substituents. Similarly, an unsaturated glucuronyl fluoride was found to be a particularly poor substrate, with kcat/Km=44 nM ?1 s?1—a very unusual result for a glycoside‐cleaving enzyme. These results are consistent with a transition state with positive charge at carbon 5 and the endocyclic oxygen, as anticipated in the hydration mechanism proposed. However, several analogues designed to take advantage of strong enzyme binding to such a transition state showed little to no inhibition. This result suggests that further work is required to understand the true nature of the transition state stabilised by this enzyme.  相似文献   

9.
The third‐generation peptide‐dendrimer B1 (AcES)8(BEA)4(K‐Amb‐Y)2BCD‐NH2 (B=branching (S)‐2,3‐diaminopropanoic acid, K=branching lysine, Amb=4‐aminomethyl‐benzoic acid) is the first synthetic model for cobalamin‐binding proteins and binds cobalamin strongly (Ka=5.0×106 M ?1) and rapidly (k2=346 M ?1 s?1) by coordination of cobalt to the cysteine residue at the dendrimer core. A structure–activity relationship study is reported concerning the role of negative charges in binding. Substituting glutamates (E) for glutamines (Q) in the outer branches of B1 to form N3 (AcQS)8(BQA)4(B‐Amb‐Y)2BCD‐NH2 leads to stronger (Ka=12.0×106 M ?1) but slower (k2=67 M ?1 s?1) cobalamin binding. CD and FTIR spectra show that the dendrimers and their cobalamin complexes exist as random‐coil structures without aggregation in solution. The hydrodynamic radii of the dendrimers determined by diffusion NMR either remains constant or slightly decreases upon binding to cobalamin; this indicates the formation of compact, presumably hydrophobically collapsed complexes.  相似文献   

10.
Mannopeptimycin, a potent drug lead, has superior activity against difficult‐to‐treat multidrug‐resistant Gram‐positive pathogens such as methicillin‐resistant Staphylococcus aureus (MRSA). (2S,3S)‐β‐Methylphenylalanine is a residue in the cyclic hexapeptide core of mannopeptimycin, but the synthesis of this residue is far from clear. We report here on the reaction order and the stereochemical course of reaction in the formation of (2S,3S)‐β‐methylphenylalanine. The reaction is executed by the enzymes MppJ and TyrB, an S‐adenosyl methionine (SAM)‐dependent methyltransferase and an (S)‐aromatic‐amino‐acid aminotransferase, respectively. Phenylpyruvic acid is methylated by MppJ at its benzylic position at the expense of one equivalent of SAM. The resulting β‐methyl phenylpyruvic acid is then converted to (2S,3S)‐β‐methylphenylalanine by TyrB. MppJ was further determined to be regioselective and stereoselective in its catalysis of the formation of (3S)‐β‐methylphenylpyruvic acid. The binding constant (KD) of MppJ versus SAM is 26 μM . The kinetic constants with respect to kcat Ppy and KM Ppy, and kcat SAM and KM SAM are 0.8 s?1 and 2.5 mM , and 8.15 s?1 and 0.014 mM , respectively. These results suggest SAM has higher binding affinity for MppJ than Ppy, and the C? C bond formation in βmPpy might be the rate‐limiting step, as opposed to the C? S bond breakage in SAM.  相似文献   

11.
Specific inhibition of the copper‐containing peptidylglycine α‐hydroxylating monooxygenase (PHM), which catalyzes the post‐translational modification of peptides involved in carcinogenesis and tumor progression, constitutes a new approach for combating cancer. We carried out a structure–activity study of new compounds derived from a well‐known PHM substrate analogue, the olefinic compound 4‐phenyl‐3‐butenoic acid (PBA). We designed, synthesized, and tested various PBA derivatives both in vitro and in silico. We show that it is possible to increase PBA affinity for PHM by appropriate functionalization of its aromatic nucleus. Compound 2 d , for example, bears a meta‐benzyloxy substituent, and exhibits better inhibition features (Ki=3.9 μM , kinact/Ki=427 M ?1 s?1) than the parent PBA (Ki=19 μM , kinact/Ki=82 M ?1 s?1). Docking calculations also suggest two different binding modes for PBA derivatives; these results will aid in the development of further PHM inhibitors with improved features.  相似文献   

12.
Injected chemical flooding systems with high salinity tolerance and fast‐dissolving performance are specially required for enhancing oil recovery in offshore oilfields. In this work, a new type of viscoelastic‐surfactant (VES) solution, which meets these criteria, was prepared by simply mixing the zwitterionic surfactant N‐hexadecyl‐N,N‐dimethyl‐3‐ammonio‐1‐propane sulfonate (HDPS) or N‐octyldecyl‐N,N‐dimethyl‐3‐ammonio‐1‐propane sulfonate (ODPS) with anionic surfactants such as sodium dodecyl sulfate (SDS). Various properties of the surfactant system, including viscoelasticity, dissolution properties, reduction of oil/water interfacial tension (IFT), and oil‐displacement efficiency of the mixed surfactant system, have been studied systematically. A rheology study proves that at high salinity, 0.73 wt.% HDPS/SDS‐ and 0.39 wt.% ODPS/SDS‐mixed surfactant systems formed worm‐like micelles with viscosity reaching 42.3 and 23.8 mPa s at a shear rate of 6 s?1, respectively. Additionally, the HDPS/SDS and ODPS/SDS surfactant mixtures also exhibit a fast‐dissolving property (dissolution time <25 min) in brine. More importantly, those surfactant mixtures can significantly reduce the IFT of oil–water interfaces. As an example, the minimum of dynamic‐IFT (IFTmin) could reach 1.17 × 10?2 mN m?1 between the Bohai Oilfield crude oil and 0.39 wt.% ODPS/SDS solution. Another interesting finding is that polyelectrolytes such as sodium of polyepoxysuccinic acid can be used as a regulator for adjusting IFTmin to an ultralow level (<10?2 mN m?1). Taking advantage of the mobility control and reducing the oil/water IFT of those surfactant mixtures, the VES flooding demonstrates excellent oil‐displacement efficiency, which is close to that of polymer/surfactant flooding or polymer/surfactant/alkali flooding. Our work provides a new type of VES flooding system with excellent performances for chemical flooding in offshore oilfields.  相似文献   

13.
The kinetics of oxidation with ozone of 1,3,6‐naphthalenetrisulfonic acid was analysed by studying the influence of different experimental parameters such as the concentration of tert‐butyl alcohol (2‐methyl‐2‐propanol), initial concentration of the acid, pH, and temperature. The rate constant of the direct reaction at 25 °C was calculated (kD = 6.72 M ?1s?1). The constant of the free radical reaction was determined with the competitive kinetics method, using sodium 4‐chlorobenzoate as reference compound, obtaining a value of kOH = 3.7 × 109 M ?1s?1. It was demonstrated that even at very acid pH values, 80% of the 1,3,6‐naphthalenetrisulfonic acid was degraded by free radical reactions, so that the ozonation of this acid may be considered an advanced oxidation process. © 2002 Society of Chemical Industry  相似文献   

14.
The reductive amination of ketones to produce chiral amines is an important transformation in the production of pharmaceutical intermediates. Therefore, industrially applicable enzymatic methods that enable the selective synthesis of chiral amines could be very useful. Using a phenylalanine dehydrogenase scaffold devoid of amine dehydrogenase activity, a robust amine dehydrogenase has been evolved with a single two‐site library allowing for the direct production of (R)‐1‐(4‐fluorophenyl)‐propyl‐2‐amine from para‐fluorophenylacetone with a kcat value of 6.85 s−1 and a KM value of 7.75 mM for the ketone substrate. This is the first example of a highly active amine dehydrogenase capable of accepting aliphatic and benzylic ketone substrates. The stereoselectivity of the evolved amine dehydrogenase was very high (>99.8% ee) showing that high selectivity of the wild‐type phenylalanine dehydrogenase was conserved in the evolution process. When paired with glucose/glucose dehydrogenase, NADH cofactor can be effficiently regenerated and the reaction driven to over 93% conversion. The broad specificity, high selectivity, and near complete conversion render this amine dehydrogenase an attractive target for further evolution toward pharmaceutical compounds and subsequent application.  相似文献   

15.
BACKGROUND: The effect of acrylic acid neutralization on the degradation of alkoxyamine initiators for nitroxide‐mediated polymerization (NMP) was studied using styrene/acrylic acid and styrene/sodium acrylate random copolymers (20 mol% initial acrylate feed concentration) as macro‐initiators. The random copolymers were re‐initiated with fresh styrene in 1,4‐dioxane at 110 °C at SG1 mediator/BlocBuilder® unimolecular initiator ratios of 5 and 10 mol%. RESULTS: The value of kpK (kp = propagation rate constant, K = equilibrium constant) was not significantly different for styrene/acrylic acid and styrene/sodium acrylate compositions at 110 °C (kpK = 2.4 × 10?6–4.6 × 10?6 s?1) and agreed closely with that for styrene homopolymerization at the same conditions (kpK = 2.7 × 10?6–3.0 × 10?6 s?1). All random copolymers had monomodal, narrow molecular weight distributions (polydispersity index M?w/M?n = 1.10–1.22) with similar number‐average molecular weights M?n = 19.3–22.1 kg mol?1. Re‐initiation of styrene/acrylic acid random copolymers with styrene resulted in block copolymers with broader molecular weight distributions (M?w/M?n = 1.37–2.04) compared to chains re‐initiated by styrene/sodium acrylate random copolymers (M?w/M?n = 1.33). CONCLUSIONS: Acrylic acid degradation of the alkoxyamines was prevented by neutralization of acrylic acid and allowed more SG1‐terminated chains to re‐initiate the polymerization of a second styrenic block by NMP. Copyright © 2008 Society of Chemical Industry  相似文献   

16.
The self‐assembling anionic surfactant, sodium n‐dodecyl sulfonate (SDS) micelles were used as pore‐forming templating for fabricating novel superporous hydroxyethyl cellulose‐grafting‐poly(sodium acrylate)/attapulgite (HEC‐g‐PNaA/APT) hydrogels. The network characteristics, morphologies of the hydrogels and removing of SDS micelles from the final product by washing with ethanol/water (v/v, 7 : 3) procedure were determined by Fourier transform infrared spectroscopy and scanning electron microscopy, as well as by determination of swelling ratio, swelling rate, and stimuli response to salts and pHs. The results showed that the added‐SDS concentration significantly affected the morphologies and pore structure of the hydrogel, and 2 mM SDS facilitates to form a homogeneous and well‐defined pore structure in the gel network to extremely improve the swelling ratio and swelling rate. The 2 mM SDS‐added superporous HEC‐based hydrogel not only had highest equilibrium swelling ratio (Qeq, 1118, 102 g g?1 in distilled water and 0.9 wt % NaCl solution), rapid swelling rate (kis, 5.2840 g g s?1), also showed multistimulus responses to salts and pHs, which may allow its applications in several areas such as adsorption, separation and biomedical materials. © 2015 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 42027.  相似文献   

17.
An unexpected, redox‐neutral C?C bond isomerization of a γ‐butyrolactone bearing an exo‐methylene unit to the thermodynamically more favoured endo isomer (kcat=0.076 s?1) catalysed by flavoproteins from the Old Yellow Enzyme family was discovered. Theoretical calculations and kinetic data support a mechanism through which the isomerization proceeds through FMN‐mediated hydride addition onto exo‐Cβ, followed by hydride abstraction from endo‐Cβ′, which is in line with the well‐established C?C bond bioreduction of OYEs. This new isomerase activity enriches the catalytic versatility of ene‐reductases.  相似文献   

18.
Zinc‐dependent alcohol dehydrogenases (ADHs) are a class of enzymes applied in different biocatalytic processes ranging from lab to industrial scale. However, one drawback is the limited substrate range, necessitating a whole array of different ADHs for the relevant substrate classes. In this study, we investigated structural determinants of the substrate spectrum in the zinc‐dependent ADH carbonyl reductase 2 from Candida parapsilosis (CPCR2), combining methods of mutational analysis with in silico substrate docking. Assigned active site residues were genetically randomized, and the resulting mutant libraries were screened with a selection of challenging carbonyl substrates. Three variants (C57A, W116K, and L119M) with improved activities toward different substrates were detected at neighboring positions in the active site. Thus, all possible combinations of the mutations were generated and characterized for their substrate specificity, yielding several improved variants. The most interesting were a C57A variant, with a 27‐fold increase in specific activity for 4′‐acetamidoacetophenone, and the double mutant CPCR2 B16‐(C57A, L119M), with a 45‐fold improvement in the kcat?KM?1 value. The obtained variants were further investigated by in silico docking experiments. The results indicate that the mentioned residues are structural determinants of the substrate specificity of CPCR2, being major players in the definition of the active site. Comparison of these results with closely related enzymes suggests that these might even be transferred to other ADHs.  相似文献   

19.
Cathepsin C is a papain‐like cysteine protease with dipeptidyl aminopeptidase activity that is thought to activate various granule‐associated serine proteases. Its exopeptidase activity is structurally explained by the so‐called exclusion domain, which blocks the active‐site cleft beyond the S2 site and, with its Asp 1 residue, provides an anchoring point for the N terminus of peptide and protein substrates. Here, the hydrazide of (2S,3S)‐trans‐epoxysuccinyl‐L ‐leucylamido‐3‐methylbutane (E‐64c) (k2/Ki=140±5 M ?1 s?1) is demonstrated to be a lead structure for the development of irreversible cathepsin C inhibitors. The distal amino group of the hydrazide moiety addresses the acidic Asp 1 residue at the entrance of the S2 pocket by hydrogen bonding while also occupying the flat hydrophobic S1′–S2′ area with its leucine‐isoamylamide moiety. Furthermore, structure–activity relationship studies revealed that functionalization of this distal amino group with alkyl residues can be used to occupy the conserved hydrophobic S2 pocket. In particular, the n‐butyl derivative was identified as the most potent inhibitor of the series (k2/Ki=56 000±1700 M ?1 s?1).  相似文献   

20.
The aggregation behavior of a di‐ and tri‐block copolymers of type PEO‐PBO, PEO‐PBO‐PEO, surface‐active ionic liquid (SAIL) of type 4‐dodecyl‐4‐methylmorpholinium chloride [C12mmor][Cl], and 1‐dodecyl‐1‐methylpyrrolidinium chloride [C12mpyrr][Cl]) in water as well as in 10 mM of a poorly water soluble dexamethasone (dex) aqueous solution was studied by determining the critical micelle concentrations using drug solubilization, surface tension, and isothermal titration calorimetry (ITC) methods. ITC measurements were also made on solutions prepared by mixing the micellar aqueous solutions of copolymers and simple aqueous solutions of SAIL across the mole fractions at three different temperatures (298.15, 308.15, and 318.15 K). The thermodynamic parameters, namely Gibbs free energy (ΔGm), enthalpy (ΔHm), and entropy (ΔSm), of micellization were calculated, and it was observed that the negative ΔGm and positive ΔSm for the mixture solutions increase with the increase in mole fraction of SAIL. Otherwise, the micellization is reported to be a spontaneous and highly entropy‐driven process. The dex‐solubilized micellar solutions were mixed with agar to obtain standing gels. The gel samples were dry‐cast into thin films, and the release of dex from films by simple dilution was monitored by UV measurements. The drug release data was fitted to several mechanistic models, and it was inferred that the release mechanism for dex from thin films is non‐Fickian for mixtures and Fickian in copolymer or SAIL micellar aqueous solutions. The transport of dex is diffusion‐controlled with diffusivities of 5.8–12 × 10?11 m2 s?1 for copolymer micelles, 5–11 × 10?11 m2 s?1 for micelles of SAIL, and 3–14 × 10?11 m2 s?1 for the mixed micelles of copolymer and SAIL in aqueous media.  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