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.
A molecular dynamics simulation was carried out involving a paclitaxel molecule, 987 peptoad molecules, and 35 938 water molecules (conditions shown experimentally to effect paclitaxel solubilization in water). The peptoads are shown to form large clumps, the centers of which are dry and thus favorable to hydrogen bonding between paclitaxel and peptoads. Hydrogen-bonding equilibrium among the peptoads themselves in the developing clumps is achieved in 2 ns. The number and position of hydrogen bonds between the paclitaxel and peptoads fluctuate randomly from two to six within a 2-5 ns time frame. Hydrophobic association between the peptoad chains and the apolar paclitaxel groups does not seem to be an important element of the solubilization. Instead, the hydrophobic chains of the peptoads encasing the paclitaxel extend outward into the dry interior of the peptoad clump where other chains in the clump are located. One hopes that studies such as this will ultimately allow rational predictions when designing new and specific drug solubilizers.  相似文献   

2.
Paclitaxel (Taxol), a diterpenoid isolated from Taxus brevifolia, is effective against several murine tumors, and is one of the most exciting anticancer molecules currently available. Due to its low solubility in water, it is clinically administered with polyethoxylated castor oil (Cremophor EL), which causes serious side effects. Inclusion of paclitaxel in solid lipid nanoparticles (SLNs) has proved to be a good approach to eliminate the need for Cremophor EL and improve the drug's antitumor efficacy. This paper describes the development of two types of long-circulating SLNs as colloidal carriers for paclitaxel. SLNs are constituted mainly of bioacceptable and biodegradable lipids. In vitro release kinetics showed that the release was very slow, the release of paclitaxel from F68-SLN is linear, and the release of paclitaxel from Brij78-SLN followed the Weibull equation. Pharmacokinetics was evaluated in KM mice after injection of paclitaxel formulated in Cremophor EL or in Brij78-SLN and F68-SLN. Encapsulation of paclitaxel in both SLNs produced marked differences compared with the free drug pharmacokinetics. F68-SLN and Brij78-SLN are long-circulating (t 1/2 beta, 10.06 and 4.88 h, respectively) compared with paclitaxel injection (t 1/2 beta, 1.36 h).  相似文献   

3.
Introduction Paclitaxel (as Taxol) is a kind of diterpenoid natural product[1] extracted from Chinese yew. It has been reported to have high anti-tumor effects, such as the activity against oophorama, mammary cancer, encephaloma, cervical carcinoma, and non- small-cell lung carcinoma[2-5]. One of the major problems of paclitaxel applied to therapy is its extremely low solubility in water[6]. In addition, paclitaxel is administered as a slow infusion in a vehicle consisting of Cremophor EL (polyoxyethylated castor oil ). However, Cremophor has been observed to cause severe, occasionally fatal hypersensitivity reactions in animals and humans[7].Therefore, paclitaxel therapy includes a prophylactic regimen of antihistamines and corticosteroids[8] , along with a prolonged infusion time to reduce the severity and incidence of hypersensitivity reactions. Because of the reasons mentioned above, currently its preparation needs to be improved further[9].  相似文献   

4.
A severe limitation for cancer therapy is the poor water solubility of many important therapeutic anticancer drugs. The development of novel delivery systems is therefore currently ongoing. We propose the use of β-cyclodextrin based nanosponges to deliver paclitaxel as an alternative to classical formulation in Cremophor EL. They are solid nanoparticles with mean diameter lower than 500?nm and spherical shape. Nanosponges show a safe profile being non-hemolytic and non cytotoxic. Nanosponges dissolved and encapsulated paclitaxel up to 2?mg/ml. The paclitaxel-loaded nanosponges formed a water stable colloidal system avoiding the recrystallization of paclitaxel. The in vitro release studies showed an almost complete release in 2?h without initial burst effect. Our study demonstrates that delivery of paclitaxel via nanosponges increased the amount of paclitaxel entering cancer cells and lowers paclitaxel IC50, therefore enhancing its pharmacological effect. β-Cyclodextrin based nanosponges can therefore be considered an alternative system to solubilize and deliver the paclitaxel.  相似文献   

5.
Treatment of 4,5-bis[4-(methoxycarbonyl)phenoxy]phthalonitrile and 4,5-bis[3,5-bis(methoxycarbonyl)phenoxy]phthalonitrile with an excess of 1,3-diiminoisoindoline in the presence of Zn(OAc)(2)·2H(2)O and 1,8-diazabicyclo[5.4.0]undec-7-ene in triethylene glycol monomethyl ether or polyethylene glycol monomethyl ether (with an average molecular weight of 550) led to "3 + 1" mixed cyclisation and transesterification in one pot, affording the corresponding di-β-substituted zinc(II) phthalocyanines in 7-23% yield. As shown by absorption spectroscopy, these compounds were essentially non-aggregated in N,N-dimethylformamide and could generate singlet oxygen effectively. The singlet oxygen quantum yields (Φ(Δ) = 0.53-0.57) were comparable with that of the unsubstituted zinc(II) phthalocyanine (Φ(Δ) = 0.56). These compounds in Cremophor EL emulsions also exhibited photocytotoxicity against HT29 human colorectal adenocarcinoma and HepG2 human hepatocarcinoma cells with IC(50) values in the range of 0.25-3.72 μM. The analogue with four triethylene glycol chains was the most potent photosensitiser and localised preferentially in the mitochondria of HT29 cells. The bis(polyethylene glycol)-counterpart could form surfactant-free nanoparticles both in water and in the culture medium. The hydrodynamic radii, as determined by dynamic laser light scattering, ranged from 6.3 to 79.8 nm depending on the preparation methods and conditions. The photocytotoxicity of these nanoparticles (IC(50) = 0.43-0.56 μM) was comparable with that of the Cremophor EL-formulated system (IC(50) = 0.34 μM).  相似文献   

6.
Dichlorosilicon phthalocyanine (Cl2SiPc) and bis(tri-n-hexylsiloxy) silicon phthalocyanine (HexSiPc) have been evaluated in vitro as potential photosensitizers for photodynamic therapy (PDT) against the human amelanotic melanoma cell line M6. Each photosensitizer is dissolved in a solvent-PBS mixture, or entrapped in egg-yolk lecithin liposomes or in Cremophor EL micelles. The cells are incubated for 1 h with the sensitizer and then irradiated for 20 min, 1 h or 2 h (lambda > 480 nm, 10 mW cm-2). The photocytotoxic effect is dependent on the photosensitizer concentration and the light dose. Higher phototoxicity is observed after an irradiation of 2 h: treatment with a solution of photosensitizer (2 x 10(-9) M) leads to 10% (HexSiPc in egg-yolk lecithin liposomes) or 20% (Cl2SiPc in DMF-PBS solution) cell viability. After 1 h incubation and 20 min of light exposure, the photodynamic effect is connected with the type of delivery system used. For HexSiPc, lower cell viability is found when this photosensitizer is entrapped in egg-yolk lecithin instead of solvent-PBS or for Cremophor EL micelles with Cl2SiPc. Liposome-delivered HexSiPc leads to lipid damage in M6 cells, illustrated by an increase of thiobarbituric acid-reacting substances (TBARs), but the change is not significant with Cremophor EL. The same is observed for the antioxidative defences after photodynamic stress. The cells irradiated with HexSiPc entrapped in liposomes display an increase of superoxide dismutase (SOD) activity and a decrease of glutathione (GSH) level, glutathione peroxidase (GSHPx) and catalase (Cat) activities.  相似文献   

7.
A strategy is developed for the identification of isocephalomannine in the presence of alkali metal ion adducts and other cephalomannine isomers in a paclitaxel active pharmaceutical ingredient. Intact molecular ion analyses and a sub-structural study have been performed for the differentiation of isocephalomannine (2-debenzoylpaclitaxel-2-pentenoate) from cephalomannine and 7-epi-cephalomannine. A comparative study of the cephalomannine isomers was carried out using molecular ions (MS) and fragmentation patterns (MS/MS) for sub-structural analysis. An attempt has been made to identify isocephalomannine in Cremophor(R) EL formulations.  相似文献   

8.
W/O nano-emulsion formation by a low-energy emulsification method is described for the first time. The nano-emulsions have been formed in water/mixed Cremophor EL:Cremophor WO7 surfactant/isopropyl myristate systems at Cremophor EL:Cremophor WO7 ratios between 1:2 and 1:9, by slow addition of isopropyl myristate to surfactant/water mixtures. Phase behaviour studies have showed that the compositions giving rise to W/O nano-emulsions belong to multiphase regions, one of the phases being a lamellar liquid crystalline phase. The droplet size of the nano-emulsions at a fixed oil concentration of 85% and mixed surfactants/water ratio of 70/30 ranged from 60 to 160 nm as Cremophor EL:Cremophor WO7 ratio increased from 1:8 to 1:2. These nano-emulsions showed high kinetic stability. No phase separation was observed during 5 months in nano-emulsions of the water/Cremophor EL:Cremophor WO7 1:8/isopropyl myristate system with 85% oil concentration, although droplet size experienced an increase with time.  相似文献   

9.
Two polyisoprene-polyethyleneoxide diblock copolymers with different block length ratios adsorbed to the water surface were investigated by multiple angle of incidence ellipsometry, evanescent wave light scattering, and surface tension experiments. In a semidilute interfacial regime, the transition from a two-dimensional to a "mushroom" regime, in which polymer chains form loops and tails in the subphase, was discussed. A diffusion mechanism parallel to the interface was probed by evanescent wave dynamic light scattering. At intermediate concentrations, the interfacial diffusion coefficient D(∥) scales with the surface concentration Γ, as D(∥) ~ Γ(0.77) in agreement with the scaling observed for polymer solutions in a semidilute regime. At relatively high concentrations a decreasing of D(∥) is discussed in terms of increasing friction due to interactions between polyisoprene chains.  相似文献   

10.
11.
The formation of complexes between the cationic surfactant dodecyl trimethylammonium bromide (DTAB) and the comb-type anionic polyelectrolytes poly(sodium acrylate-co-sodium 2-acrylamido-2-methylpropane sulfonate)-g-poly(N,N-dimethylacrylamide) (P(NaA-co-NaAMPS)-g-PDMAMx) was investigated in dilute aqueous solutions by means of turbidimetry, pyrene fluorescence probing, viscometry, z-potential measurements, and dynamic light scattering. The comb-type copolymers consist of an anionic copolymer backbone, P(NaA-co-NaAMPS), containing 84 mol % NaAMPS units, while the weight percentage, x, of the PDMAM side chains varies from x = 12% (w:w) up to x = 58% (w:w). It was found that, contrary to the water-insoluble complexes formed between the linear polyelectrolyte P(NaA-co-NaAMPS) and DTAB, the solubility in water of the complexes formed between the comb-type copolymers and DTAB is significantly improved with increasing x. The complexation process starts at the same critical aggregation concentration (about 2 orders of magnitude lower than the critical micelle concentration of DTAB), regardless of x, and it is accompanied by charge neutralization and appearance of hydrophobic microdomains. Both effects lead to the substantial collapse of the polyelectrolyte chain upon addition of DTAB. However, the complexes of the comb-type copolymers with DTAB are stabilized in water as nanoparticles, and probably consisted of a water-insoluble core (the polyelectrolyte/surfactant complex), protected by a hydrophilic nonionic PDMAM corona. The size of the nanoparticles varies from approximately 35 nm up to approximately 120 nm, depending on x.  相似文献   

12.
[reaction: see text] Synthetic strategies for preparing N,N'-bridged oligomeric guanidines bearing peptide side chains both on solid support and in solution are presented. Monomers are prepared from common alpha-amino acids and therefore contain conventionally protected peptide side chains. The side chains include alkyl, aromatic, hydroxyl, amino, carboxylic acid, and amide functional groups. Oligomer elongation utilizes acid-sensitive sulfonyl activated thiourea through the formation of carbodiimide intermediate. With proper preparation of monomers, synthesis of oligomer can be performed in two directions (equivalent to N to C terminal or C to N terminal in a peptide sequence) with excellent efficiency.  相似文献   

13.
Colloid-polymer (CP) mixtures extend between two limiting cases, the colloid limit with the polymer coil size small compared to the colloid radius Rcol and the protein limit with the colloidal particles much smaller in size than the radius of gyration of the polymer chains Rg. In the present work, model systems are developed for the protein limit. The colloid-solvent pairs are optimized in terms of their isorefractivity in order to facilitate the characterization of large polystyrene chains in suspensions of small colloids. The degree of isorefractivity of colloidal particles was successfully evaluated in terms of a reduced scattering intensity. Two polystyrene samples with radii of gyration of Rg = 96 nm and Rg = 78 nm, respectively, are used. The radii of the colloidal particles are close to Rcol = 12 nm, leading to size ratios of Rg/Rcol = 8 and Rg/Rcol = 6.5. Four colloid solvent systems were found to be suitable for polymer characterization by light scattering, one based on silica particles and three systems with acrylate particles. The present investigation is focused on the three acrylate systems: poly(methyl methacrylate) in ethyl benzoate (ETB) at 7 degrees C, poly(ethyl methacrylate) in toluene at 7 degrees C and poly(ethyl methacrylate) in ETB at 40 degrees C. Characterization of PS chains is for the first time performed in colloid concentrations up to 2.5% by weight. In all cases, the size and shape of the polymer chains remain unchanged. A slight mismatch of the colloid scattering or a limited colloid solubility prevented investigation of PS chains at higher colloid concentration.  相似文献   

14.
Lentinan, a beta-(1-->3)-D-glucan, was isolated from Lentinus edodes by using an improved extraction and purification method to show good water solubility and high yield. The results from 13C NMR, size-exclusion chromatography combined with multiangle laser light scattering (SEC-MALLS), dynamic light scattering (DLS), and optical rotation revealed that lentinan existed in a triple-helical conformation in the aqueous solution at 25 degrees C, whereas the thermally induced conformation transition from triple helix to single flexible chains occurred at elevated temperatures. The dependences of the weight-average molecular weight (Mw), radius of gyration (z1/2), hydrodynamic radius (Rh), intrinsic viscosity ([eta]), and specific optical rotation of lentinan on temperature in 0.9% NaCl aqueous solution showed an abrupt drop at 130-145 degrees C. It was confirmed that the conformation transitions from triple strand to single chain and from extended chains to winding chains for lentinan were completed rapidly at 130-145 degrees C, as a result of the simultaneous destruction of the intra- and intermolecular hydrogen bonds in lentinan. The thermally induced conformational transition was irreversible. The results from atomic force microscopy (AFM) and DLS demonstrated the existence of intrachain entanglement for the triple-helical chains, leading to the wormlike linear, circular, and crossover species for lentinan having high Mw (1.71x10(6)) in aqueous solution at 25 degrees C.  相似文献   

15.
The pseudodipeptide, (S)‐N‐iso­propyl {[N‐(pivaloyl)­pyrrol­idin‐2‐yl]­methyl­amino­oxy}acet­amide, C15H29N3O3, adopts a global extended conformation with the hydroxy­l­amine group in the g+/g? structure. The C‐terminal amide NH interacts intramolecularly with the hydroxy­lamine O atom. Both NH bonds of each mol­ecule are hydrogen bonded to the C‐­terminal amide carbonyl of a neighbouring mol­ecule.  相似文献   

16.
Using 1-anilino,8-naphthalenesulfonic acid (ANS) as a probe, we examined properties of micelles of Cremophor EL, an amphipathic agent which can solubilize hydrophobic photosensitizing agents and promote their distribution to plasma lipoprotein. In aqueous solution, Cremophor micelles persisted for several hours after dilution below the critical micellar concentration (CMC). After equilibrium was reached, we found a CMC of 0.009% (wt/vol). Fluorescence data suggest that the micellar environment of ANS binding has a dielectric constant of approximately 27. Cremophor also reverses examples of multi-drug resistance associated with impaired accumulation of anti-tumor agents, e.g. daunorubicin. Although the latter drug is relatively hydrophilic, fluorescence spectroscopy and anisotropy studies indicate an association with Cremophor. Moreover, resistance reversal occurred only at Cremophor concentrations above the CMC.  相似文献   

17.
Complexes containing pyridine-2-carboxaldehyde (pyca) ligand acting as κ(2)-(N,O) chelates in [MX(CO)(3)(pyca)] (M = Mn, Re; X = Cl, Br), or [MoX(methallyl)(CO)(2)(pyca)] (X = Cl, Br), are good precursors for iminopyridine complexes derived from amino esters and peptides of formula [MX(CO)(3)(py-2-C(H)═NCHX-COOY)] or [MoX(methallyl)(CO)(2)(py-2-C(H)═NCHX-COOY)], via Schiff condensation of the aldehyde function of pyca with the terminal NH(2) group of the amino ester or peptide. X-ray determinations confirm the structures and show that in solid phase the peptide chains assemble through H-bonds adopting different patterns which depend on the geometry of the metal-ligand fragments. The H-bonding patterns have been analyzed in detail and described by using graph set methods. In most cases, Mo complexes show intramolecular arrangement involving the halogen (Cl or Br) and an NH group of the side chain. For the Mn and Re complexes, the peptide side arms form infinite chains, helices, and rings. In many cases, the terminal carboxylic O-H function is engaged in a "terminal" H-bond with a polar molecule of solvent (THF or acetone), instead of forming the usual head-to-head arrangement found in simple carboxylic acids. For the longer tripeptide Gly-Gly-Gly, a discrete, dimeric association is observed, in which the peptide chains show antiparallel arrangement with a complementary disposition of the internal N-H and C═O functions. DOSY experiments in solution show significant changes in the diffusion rates upon addition of OPBu(3), which indicate H-bonding interaction of OPBu(3) with the peptide hydrogens.  相似文献   

18.
Sugar-lipid hybrids of the type CnCm were prepared by coupling an alkane chain (Cn) with a maltooligosaccharide (Gm) over an amide linkage. Coupling was performed with maltobionolactone (G2) and n-alkylamine chains Cn withn=8,10,12,14,16, i.e. variation of the hydrophobic part of the molecule, and with hexadecylamine (C16) and different maltooligosaccharides (Gm, m=2,3,4,6). The solution properties of the various products were studied by means of static and dynamic light scattering (LS) and by electron-microscopy (EM).The results may be summarized as follows: If the alkane chain is shorter thann=14, small spherical micelles with a radius of about 3 nm are observed. In time these micelles aggregate further to form increasingly larger spherical clusters which eventually precipitate. Long rod-like micelles form whenn 14. Contour length and chain stiffness were determined by applying theories of semiflexible chains. A qualitative confirmation of the light scattering results, i.e., micelle size and shape, was obtained from electron microscopy.  相似文献   

19.
Fubin Chen 《Tetrahedron》2008,64(40):9585-9591
A series of proline-based dipeptide organocatalysts with two amide units (1-16) have been developed and evaluated in the direct catalytic asymmetric aldol reactions of aldehydes with cyclohexanone. These catalysts showed good solubility in organic solvents compared with their corresponding carboxyl terminal dipeptides. The robust amide bond formation allowed structural modifications and fine tuning of catalyst properties by varying the stereo and electronic effects of the terminal amide to affect the ability of hydrogen bonding formation between the catalysts and the substrates. The reactions proceeded smoothly in high yields (up to 99%), enantioselectivities (up to 98% ee) and anti-diastereoselectivities (up to 99:1) in the presence of bifunctional organocatalyst 4 under the optimal reaction conditions.  相似文献   

20.
以天然可生物降解的壳聚糖为原料, 通过在壳聚糖6位羟基上引入羧甲基, 在2位氨基上引入疏水烷基链, 制得N-辛基-O,N-羧甲基壳聚糖(OCC)两亲性衍生物, 分别用FTIR、1H NMR和元素分析等技术对其结构进行表征, 用广角X射线衍射(WAXD)和差示扫描量热法(DSC)对其物理性质进行分析, 并考察其在各种溶剂中的溶解性能及其对难溶性药物的增溶能力. 所制备的OCC羧甲基取代度为115.9%, 取代主要发生在6位羟基上; 辛基取代度58.0%, 取代主要发生在2位氨基上; 与壳聚糖相比, OCC分子间/内氢键作用减弱; OCC在常用的有机溶剂中不溶, 但在水中溶解度增加, 能够形成具有淡蓝色乳光的纳米胶体溶液, 对难溶性抗肿瘤药物紫杉醇具有优越的增溶能力, 使紫杉醇在水中的溶解度提高近500倍, 载药量为34.6%(质量分数), 包封率为89.9%. OCC是潜在的优良的难溶性药物增溶载体材料.  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