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 15 毫秒
1.
In this work, we simulate the microphase separation of aqueous and non-aqueous solutions of diblock copolymer poly(styrene)-b-poly(ethyleneglycol) under different architectures (linear and linear–dendritic) by dissipative particle dynamics. The observed morphologies in water where poly(ethyleneglycol) (PEG) block is soluble are as follows: (1) at low concentrations spherical micelles, cylinders and bicontinuous structures are formed in dendritic structures and spheres, cylinders and perforated lamellas in linear structure. The architectures simulated at low–moderate concentrations show an evolution sphere → cylinder → bicontinuous or perforated lamellas as the concentration is increased. (2) At high concentrated solutions rich defect structures of the sponge type are formed. In a non-aqueous non-polar solution such as cyclohexane, which is a good solvent for the polystyrene block, the formation of well-defined aggregates at low concentrations is not observed; however, irregular structures are achieved in concentrated solutions. We compare these results with a polymeric chimera consisting of a mixture of linear poly(styrene) homopolymer and PEG homopolymer in the linear, G1 or G2 dendritic configurations. Our simulations are in agreement with the experimentally observed structures of these polymers.  相似文献   

2.
UDP-glucose dehydrogenase: substrate binding stoichiometry and affinity   总被引:3,自引:0,他引:3  
Precise structural parameters of polyribonucleotides single stranded helices are determined as well as those of double stranded helices of poly 2′-O-methyl A and of poly A at neutral and acid pH. Infrared linear dichroism investigations indicate the similarity of the conformation of the sugar-phosphate backbone of these single and double stranded helices. The angles of the phosphate group for single stranded helix at neutral pH is found to be oriented at 48° for the 02P02 bisector and at about 65° for the 02–03 line to the helix axis. Similar values were found for double stranded poly A helix at acid pH. These structural parameters obtained for the first time on single stranded polynucleotide helices are proposed to be valid for other similar helical chains such as poly A segments of nuclear or messenger RNA and single stranded CCA acceptor end of transfer RNA.  相似文献   

3.
ABA triblock copolymers [A = 2-(diisopropylamino)ethyl methacrylate), DPA or 2-(diethylamino)ethyl methacrylate), DEA; B = 2-methacryloyloxyethyl phosphorylcholine, MPC] prepared using atom transfer radical polymerization dissolve in acidic solution but form biocompatible free-standing gels at around neutral pH in moderately concentrated aqueous solution (above approximately 10 w/v % copolymer). Proton NMR studies indicate that physical gelation occurs because the deprotonated outer DPA (or DEA) blocks become hydrophobic, which leads to attractive interactions between the chains: addition of acid leads to immediate dissolution of the micellar gel. Release studies using dipyridamole as a model hydrophobic drug indicate that sustained release profiles can be obtained from these gels under physiologically relevant conditions. More concentrated DPA-MPC-DPA gels give slower release profiles, as expected. At lower pH, fast, triggered release can also be achieved, because gel dissolution occurs under these conditions. Furthermore, the nature of the outer block also plays a role; the more hydrophobic DPA-MPC-DPA triblock gels are formed at lower copolymer concentrations and retain the drug longer than the DEA-MPC-DEA triblock gels.  相似文献   

4.
Liang L  Yao P  Jiang M 《Biomacromolecules》2005,6(5):2748-2755
Apocytochrome c interacts with two copolymers: poly(isobutylene-alt-maleic acid) (PIMA) and poly(1-tetradecene-alt-maleic acid) (PTMA). The interaction leads to apocytochrome c, a conformational change from random coil to alpha-helical structure. The alpha-helix content is influenced by the copolymer concentration, the length of alkyl chain of the copolymers, and pH of the medium. The electrostatic attraction between the copolymer and protein is an indispensable factor for the folding of the protein at acid pH. The hydrophobic interaction is an important factor over the entire pH range, especially when both the copolymer and protein carry negative charges at alkaline pH. The electrostatic and hydrophobic attractions between the copolymer and protein exclude water molecules, promoting the formation of hydrogen bonds within the helical structure. On the other hand, the hydrogen bonds formed between the ionized carboxyl of the copolymer and the amide of the protein partly restrain the formation of hydrogen bonds within the helical structure when the copolymer concentration is higher at pH 6.5 and 10.5.  相似文献   

5.
Statistical copolymers were prepared from N-carboxyanhydrides of L -valine and γ-benzyl-L -glutamate in dioxan with triethylamine as an initiator. The copolymerization conversion was determined by ir spectroscopy, the copolymer composition by amino acid analysis, and the molecular weights by light scattering. The monomer reactivity ratios were found to be rVal = 0.14 and rGlu(OBzl) = 6.4. High-molecular-weight copolymers are formed even at low conversions. The content of β-structure in the copolymers was estimated from the ir spectra in copolymerization mixtures. The sequence-length distribution of L -valine and γ-benzyl-L -glutamate copolymers was calculated and its dependence on copolymerization conversion is discussed. Relations between the sequence-length distribution and the content of β-structure were studied. It was found that the content of β-structure in samples with the same composition is different for low- and high-conversion copolymers. The formation of β-structure in copolymers in the copolymerization mixture requires a certain minimal sequence length, which has been found to be about 6 valine units.  相似文献   

6.
An alpha-helical coiled-coil structure is one of the basic structural units in proteins. Hydrophilic residues at the hydrophobic positions in the coiled-coil structure play important roles in structures and functions of natural proteins. We reported here a peptide that formed a triple stranded alpha-helical coiled-coil showing the pH-dependent structural change. The peptide was designed to have two His residues at the hydrophobic positions of the center of the coiled-coil structure. The peptide folded into a triple stranded coiled-coil at neutral pH, while it unfolded at acidic pH. This construct is useful to create a protein that the structure or function is controlled by pH.  相似文献   

7.
Temperature- and pH-sensitive random copolymers of N-isopropylacrylamide (NIPAAm) and propylacrylic acid (PAA) were prepared using the reversible addition fragmentation chain transfer (RAFT) polymerization method. The lower critical solution temperatures (LCSTs) (or phase separation temperatures) of the NIPAAm-co-PAA copolymer solutions were measured by the cloud-point method. At slightly acidic conditions, the LCST decreased with increase in PAA content, which suggests that the hydrophobic propyl group of PAA has a greater influence on the LCST than the polar carboxylic acid group at those conditions. An increase of pH led to a significant increase in LCST of the copolymers due to the ionization of the -COOH group. The LCSTs were studied as a function of copolymer composition over the pH range from 5.0 to 7.0. Because the pK(a) of the polymers can be tuned to fall close to neutral pH, these polymer compositions can be designed to have phase transitions triggered near physiological pH or at slightly acidic pH values that fall within acidic gradients found in biology. The NIPAAm-co-PAA copolymers thus display tunable properties that could make them useful in a variety of molecular switching and drug delivery applications where responses to small pH changes are relevant.  相似文献   

8.
许多抗肿瘤药物因为水溶解性差,在临床上的应用受到了很大影响。胶束可将药物包载到疏水核,可显著提高药物的水溶解性,是一种极具潜力的新型给药体系。然而胶束也面临着一系列问题,比如说需要提高其在体内的动力学稳定性和热力学稳定性等。与此同时,如果想要在单一的胶束体系上实现多种功能,需要对载体材料进行繁杂的修饰,也是一个难题。由不同嵌段聚合物、聚合物/表面活性剂自组装成的混合胶束或聚离子复合物胶束,相对于单一嵌段聚合物形成的胶束而言,物理稳定性和载药能力都得到了提高。同时,通过将具有不同官能团的聚合物制备成混合胶束,可以直接方便得到多功能复合的体系。本文对混合胶束载体系统的药剂学进展进行了综述。  相似文献   

9.
Ultrastructural analysis of the gel forming green seaweed sulfated polysaccharide ulvan revealed a spherical‐based morphology (10–18 nm diameter) more or less aggregated in aqueous solution. At pH 13 in TBAOH (tetrabutyl ammonium hydroxyde) or NaOH, ulvan formed an open gel‐like structure or a continuous film by fusion or coalescence of bead‐like structures, while in acidic pH conditions, ulvan appeared as dispersed beads. Low concentrations of sodium chloride, copper or boric acid induced the formation of aggregates. These results highlight the hydrophobic and aggregative behavior of ulvan that are discussed in regard to the peculiar gel formation and the low intrinsic viscosity of the polysaccharide in aqueous solution. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 652–664, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

10.
Cholesterol mesogen containing monomer, cholesteryl acrylamido butyrate (CAB) with the novel spacer group drawn from 4-amino butyric acid has been demonstrated to exhibit good reactivity with 2-acrylamido-2-methyl-1-propane sulfonic acid (AMPS) to yield copolymers with CAB content as high as 15 mol % hitherto not achieved. The spacer group is shown to provide the twin benefits of enhanced reactivity and solubility in water. The high pK(a) at > or =9.90 of these copolymers estimated from potentiometric studies demonstrates packing of AMPS segments as ionic clusters. The higher CAB in copolymer C provides the most densely packed nonpolar microdomains. From fluorescence quenching studies, the cross-linking provided by the cholesterol chains favoring intra- or intermolecular aggregated structures has been established. At the air/solution interface, copolymer C exhibits the most close-packed structures exhibiting "a" of 41.2 A(2)/molecule. The effect of neutralization on the adsorption characteristics is investigated.  相似文献   

11.
The purpose of this study is to develop novel colon-specific drug delivery systems with pH-sensitive swelling and drug release properties. Methacrylic-type polymeric prodrugs with different content levels of 5-amino salicylic acid (5-ASA) were synthesized by free radical copolymerization of metacrylic acid (MAA), polyethylene glycol monomethacrylate (PEGMA), and a methacrylic derivative of 5-ASA (methacryloyloxyethyl 5-amino salicylate [MOES]). The copolymers were characterized, and the drug content of the copolymers was determined. The effect of copolymer composition on the swelling behavior and hydrolytic degradation was studied in simulated gastric fluid (SGF, pH 1.2) and simulated intestinal fluid (SIF, pH 7.2). The swelling and hydrolytic behavior of the copolymers was dependent on the content of MAA groups and caused a decrease in gel swelling in SGF or an increase in gel swelling in SIF. Drug release studies showed that increasing content of MAA in the copolymer enhances the hydrolysis in SIF but has no effect in SGF. The results suggest that hydrogen-bonded complexes are formed between MAA and PEG pendant groups and that these pH-sensitive systems could be useful for preparation of a controlled-release formulation of 5-ASA.  相似文献   

12.
Block copolymers containing stimuli-responsive segments provide important new opportunities for controlling the activity and aggregation properties of protein-polymer conjugates. We have prepared a RAFT block copolymer of a biotin-terminated poly(N-isopropylacrylamide) (PNIPAAm)-b-poly(acrylic acid) (PAA). The number-average molecular weight (M(n)) of the (PNIPAAm)-b-(PAA) copolymer was determined to be 17.4 kDa (M(w)/M(n) = 1.09). The PNIPAAm block had an M(n) of 9.5 kDa and the poly(acrylic acid) (PAA) block had an M(n) of 7.9 kDa. We conjugated this block copolymer to streptavidin (SA) via the terminal biotin on the PNIPAAm block. We found that the usual aggregation and phase separation of PNIPAAm-SA conjugates that follow the thermally induced collapse and dehydration of PNIPAAm (the lower critical solution temperature (LCST) of PNIPAAm is 32 degrees C in water) is prevented through the shielding action of the PAA block. In addition, we show that the cloud point and aggregation properties (as measured by loss in light transmission) of the [(PNIPAAm)-b-(PAA)]-SA conjugate also depended on pH. At pH 7.0 and at temperatures above the LCST, the block copolymer alone was found to form particles of ca. 60 nm in diameter, while the bioconjugate exhibited very little aggregation. At pH 5.5 and 20 degrees C, the copolymer alone was found to form large aggregates (ca. 218 nm), presumably driven by hydrogen bonding between the -COOH groups of PAA with other -COOH groups and also with the -CONH- groups of PNIPAAm. In comparison, the conjugate formed much smaller particles (ca. 27 nm) at these conditions. At pH 4.0, however, large particles were formed from the conjugate both above and below the LCST (ca. 700 and 540 nm, respectively). These results demonstrate that the aggregation properties of the block copolymer-SA conjugate are very different from those of the free block copolymer, and that the outer-oriented hydrophilic block of PAA shields the intermolecular aggregation of the block copolymer-SA bioconjugate at pH values where the -COOH groups of PAA are significantly ionized.  相似文献   

13.
A gram-positive, rod-shaped anaerobe (isolate F-14) was isolated from soil. This organism was identified by cellular morphology as well as by fermentative and biochemical data as Clostridium limosum. Isolate F-14 formed ursocholic acid (UC) and 7-ketodeoxycholic acid (7-KDC) from cholic acid (CA), and ursodeoxycholic acid (UDC) and 7-ketolithocholic acid (7-KLC) from chenodeoxycholic acid (CDC) in whole cell cultures, but did not transform deoxycholic acid (DC). No hydrolysis or transformation occurred when either taurine- or glycine-conjugated bile acids were incubated with F-14. The type stain of Clostridium limosum (American Type Culture Collection 25620) did not transform bile acids. The structures of ursocholic, ursodeoxycholic, 7-ketodeoxycholic, and 7-ketolithocholic acids were verified by mass spectroscopy and by thin-layer chromatography using Komarowsky's spray reagent. The organism transformed cholic and chenodeoxycholic acids at concentrations of 20 mM and 1 mM, respectively; higher concentrations of bile acids inhibited growth. Optimal yields of ursocholic and ursodeoxycholic acids were obtained at 9-24 hr of incubation and depended upon the substrate used. Increasing yields of 7-ketodeoxycholic and 7-ketolithocholic acids, and decreasing yields of ursocholic and ursodeoxycholic acids were observed with longer periods of incubation. Culture pH changed with time and was characterized by a small initial drop (0.2-0.4 pH units) and a subsequent increase to a pH (8.1-8.2) that was above the starting pH (7.4).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Polymerizations of DL -phenylalanine NCA by block copolymers of sarcosine and DL -phenylalanine, designated by (Phe)m(Sar)n and capable of reaction at the phenylalanyl terminal, were investigated in nitrobenzene solution at 25°C. With increasing n for constant m (m = 0, 1, 2, and 5), the polymerization rate greatly increased. Previously the acceleration of the initiation reaction in the polymerization of DL -phenylalanine NCA by polysarcosine (m = 0) was reported. The present results showing the acceleration by the copolymers of sarcosine and DL -phenylalanine indicate the presence of the polymer effect in the propagation reaction as well. However, the polymer effect was most marked with polysarcosine (m = 0), and decreased with increasing m. The same polymerizations by sequential copolymers composed of ten sarcosine units and two DL -phenylalanine units were also investigated. Again with these copolymer catalysts the polymerization rate was larger than that by monomeric amines. But the polymer effect decreased sharply when the phenylalanine units take positions near the terminal amine group of the copolymer catalyst. These two deteriorating effects of the phenylalanine unit have been interpreted in terms of the decrease of the flexibility of polymer chain, caused possibly by an intramolecular hydrogen bond of the phenylalanine unit.  相似文献   

15.
We evaluated the potential for polyelectrolyte induced precipitation of antibodies to replace traditional chromatography purification. We investigated the impact of solution pH, solution ionic strength and polyelectrolyte molecular weight on the degree of precipitation using the anionic polyelectrolytes polyvinylsulfonic acid (PVS), polyacrylic acid (PAA), and polystyrenesulfonic acid (PSS). As we approached the pI of the antibody, charge neutralization of the antibody reduced the antibody–polyelectrolyte interaction, reducing antibody precipitation. At a given pH, increasing solution ionic strength prevented the ionic interaction between the polyelectrolyte and the antibody, reducing antibody precipitation. With increasing pH of precipitation, there was an increase in impurity clearance. Increasing polyelectrolyte molecular weight allowed the precipitation to be performed under conditions of higher ionic strength. PVS was selected as the preferred polyelectrolyte based on step yield following resolubilization, purification performance, as well as the nature of the precipitate. We evaluated PVS precipitation as a replacement for the initial capture step, as well as an intermediate polishing step in the purification of a humanized monoclonal antibody. PVS precipitation separated the antibody from host cell impurities such as host cell proteins (HCP) and DNA, process impurities such as leached protein A, insulin and gentamicin, as well as antibody fragments and aggregates. PVS was subsequently removed from antibody pools to <1 µg/mg using anion exchange chromatography. PVS precipitation did not impact the biological activity of the resolubilized antibody. Biotechnol. Bioeng. 2009;102: 1141–1151. © 2008 Wiley Periodicals, Inc.  相似文献   

16.
Aims: To investigate the ability of lactic acid bacteria (LAB) to convert linoleic acid (LA) and α‐linolenic acid (α‐LNA) to conjugated linoleic acid (CLA) and conjugated linolenic acid (CLNA), respectively. To assess pH and temperature influences on CLA and CLNA production by Lactobacillus sakei LMG 13558. Methods and Results: A screening of 48 LAB yielded one Lactobacillus curvatus, five Lactobacillus plantarum and four Lact. sakei strains displaying linoleate isomerase (LAI) activity. CLNA conversion percentages varied largely (1–60%). CLA conversion, occurring in three strains, was lower (2–5%). The LAI gene sequences of the ten LAI‐positive strains shared 75–99% identity with the LAI gene sequence of a Lact. plantarum AS1.555. At pH 6·2, CLA and CLNA production by Lact. sakei LMG 13558 was higher at 30°C than at 20 and 25°C. At pH 5·5 (30°C) or 37°C (pH 6·2), LA was not converted and α‐LNA only slightly converted. Conclusions: LAB show strain‐dependent LAI activity. Production of CLA and CLNA is affected by pH and temperature, as shown for Lact. sakei LMG 13558. Significance and Impact of the Study: Several LAB produce CLA and/or CLNA, as shown for Lact. sakei and Lact. curvatus for the first time. These findings offer potential for the manufacturing of fermented functional foods.  相似文献   

17.
Delftia acidovorans P4a (DSMZ 10474) was grown in mineral medium on acetic acid at pH 8.0 without an additional supply of nutrients like yeast extract or polypeptone. Using acetic acid and γ‐butyrolactone (GBL), copolymers with a 4HB content from 2–90 mol % were detected in batch experiments, depending on the ratio of the both carbon substrates. Due to the different consumption rates of the individual carbon substrates a multitude of different target mole fractions were difficult to produce by fed‐batch fermentation. Therefore, the two‐stage continuous cultivation technique was applied with two fermenters connected in series. At stage 2, the optimum PHA productivity of the bioreactor and a target 4HB content of the polymer could be precisely adjusted by the composition of the two substrates. This cultivation strategy was especially convenient when toxic substrates like acetic acid and GBL were employed. Using mixtures of acetic acid and GBL (3.5–23.5 mol % GBL), copolymers with a target mole fraction of 2.7–19 % 4HB could be produced. The PHA content was in the range of 52–60 %. The dilution rates (D) of the first and second fermenter were 0.2 h–1 and 0.06 h–1, respectively.  相似文献   

18.
A study of the interaction between poly(dG)-poly(dC) and poly(rC) demonstrates that, at neutral pH and high ionic strength, there is replacement of the dC strand by poly(rC). At acid pH, formation of a triple-stranded complex which equally may involve the replacement phenomenon is observed. There is no evidence for interaction at neutral pH between poly(dG)-poly(dC) and oligo(rC), while a three-stranded complex is formed at acid pH. These data are consistent with the studies of comparative stabilities of double stranded deoxy or ribo polymers and deoxy-ribo hybrids.  相似文献   

19.
In this work, dissipative particle dynamics simulations were performed to study the self-assembly morphologies of rod–coil block copolymer polystyrene-b-poly(ethylene glycol) (PS-b-PEG) in aqueous solution under different variables. Effect of time evolution on the self-assembly morphology of PS-b-PEG was observed first. Besides, spherical, cylindrical and lamellar structures were obtained at a range of concentrations. In addition, their self-assembly morphologies could also be regulated by the PS chain length. Our simulation results can provide deeper insight into the microstructure of rod–coil block copolymers in aqueous solution, which can be useful to guide the molecular design and experimental preparation of novel rod–coil block copolymers with controlled structures.  相似文献   

20.
Self-assembly and adhesion of DOPA-modified methacrylic triblock hydrogels   总被引:1,自引:0,他引:1  
Marine mussels anchor to a variety of surfaces by secreting liquid proteins that harden and form water-resistant bonds to a variety of surfaces. Studies have revealed that these mussel adhesive proteins contain an unusual amino acid, 3,4-dihydroxy-L-phenylalanine (DOPA), which is believed to be responsible for the cohesive and adhesive properties of these proteins. To separate the cohesive and adhesive roles of DOPA, we incorporated DOPA into the midblock of poly(methyl methacrylate)-poly(methacrylic acid)-poly(methyl methacrylate) (PMMA-PMAA-PMMA) triblock copolymers. Self-assembled hydrogels were obtained by exposing triblock copolymer solutions in dimethyl sulfoxide to water vapor. As water diffused into the solution, the hydrophobic end blocks formed aggregates that were bridged by the water-soluble midblocks. Strong hydrogels were formed with polymer weight fractions between 0.01 and 0.4 and with shear moduli between 1 and 5 kPa. The adhesive properties of the hydrogels on TiO2 surfaces were investigated by indentation with a flat-ended cylindrical punch. At pH values of 6 and 7.4, the fully protonated DOPA groups were highly adhesive to the TiO2 surfaces, giving values of approximately equal to 2 J/m2 for the interfacial fracture energy, which we believe corresponds to the cohesive fracture energy of the hydrogel. At these pH values, the DOPA groups are hydrophobic and have a tendency to aggregate, so contact times of 10 or 20 min are required for these high values of the interfacial strength to be observed. At a pH of 10, the DOPA groups were hydrophilic and highly swellable, but less adhesive gels were formed. Oxidation of DOPA groups, a process that is greatly accelerated at a pH of 10, decreased the adhesive performance of the hydrogels even further.  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