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 0 毫秒
1.
Liquid crystalline copolyesters of high molecular weight were obtained by polycondensation of aromatic diols, diacyl dichlorides, oligolactides, and poly(ethylene glycol)s. Hydrophilicity of the copolyesters was controlled by the content of ethyleneoxy moieties as verified by contact angle measurements. Copolyesters with ethyleneoxy moieties showed significant enhancement of degradability under physiological conditions in comparison to copolyester without ethyleneoxy moieties, which makes these copolyesters promising materials for bone tissue engineering as also verified by hardness testing and mechanical testing.  相似文献   

2.
Oil extracts of Ukpeliede spill samples from Niger Delta (Nigeria) were analyzed by gas chromatography. The amount of polycyclic aromatic hydrocarbons (PAHs), especially the lower-molecular-weight naphthene, fluorine, phenathrene, pyrene, and benzo[a]anthracene, decreased within the sampling intervals of 2 months and 5 months. There was a predominance of three-to-six-ring PAHs over the two-ring PAHs. There was marked disappearance of n-C8 to n-C11 hydrocarbon fractions and the acyclic isoprenoids (pristane and phytane). The depletion of these molecules within the two sampling intervals suggests the possible attenuation of hydrocarbons as a result of the environmental modification within the set interval.  相似文献   

3.
The anaerobic degradability of natural and synthetic polyesters is investigated applying microbial consortia (3 sludges, 1 sediment) as well as individual strains isolated for this purpose. In contrast to aerobic conditions, the natural homopolyester polyhydroxybutyrate (PHB) degrades faster than the copolyester poly(hydroxybutyrate-co-hydroxyvalerate) (PHBV). For the synthetic polyester poly(epsilon-caroplacton) (PCL), microbial degradation in the absence of oxygen could be clearly demonstrated; however, the degradation rate is significantly lower than for PHB and PHBV. Other synthetic polyesters such as poly(trimethylene adipate) (SP3/6), poly(tetramethylene adipate) (SP4/6), and aliphatic-aromatic copolyesters from 1,4-butanediol, terephthalic acid, and adipic acid (BTA-copolymers) exhibit only very low anaerobic microbial susceptibility. A copolyester with high amount of terephthalic acid (BTA 40:60) resisted the anaerobic breakdown even under thermophilic conditions and/or when blended with starch. For the synthetic polymers, a number of individual anaerobic strain could be isolated which are able to depolymerize the polymers and selected strains where identified as new species of the genus Clostridium or Propionispora. Their distinguished degradation patterns point to the involvement of different degrading enzymes which are specialized to depolymerize either the natural polyhydroxyalkanoates (e.g., PHB), the synthetic polyester PCL, or other synthetic aliphatic polyesters such as SP3/6. It can be supposed that these enzymes exhibit comparable characteristics as those described to be responsible for aerobic polyester degradation (lipases, cutinases, and PHB-depolymerases).  相似文献   

4.
Amphiphilic graft copolymers consisting of poly(gamma-glutamic acid) (gamma-PGA) as the hydrophilic backbone and L-phenylalanine ethylester (L-PAE) as the hydrophobic side chain were synthesized by grafting L-PAE to gamma-PGA. The nanoparticles were prepared by a precipitation method, and about 200 nm-sized nanoparticles were obtained due to their amphiphilic properties. The hydrolytic and enzymatic degradation of these gamma-PGA nanoparticles was studied by gel permeation chromatography (GPC), scanning electron microscopy (SEM), dynamic light scattering (DLS) and (1)H NMR measurements. The hydrolysis ratio of gamma-PGA and these hydrophobic derivatives was found to decrease upon increasing the hydrophobicity of the gamma-PGA derivates. The pH had an effect on the hydrolytic degradation of the polymer. The hydrolysis of the polymer could be accelerated by alkaline conditions. The degradation of the gamma-PGA backbone by gamma-glutamyl transpeptidase (gamma-GTP) resulted in a dramatic change in nanoparticle morphology. With increasing time, the gamma-PGA nanoparticles began to decrease in size and finally disappeared completely. Moreover, the gamma-PGA nanoparticles were degraded by four different enzymes (Pronase E, protease, cathepsin B and lipase) with different degradation patterns. The enzymatic degradation of the nanoparticles occurred via the hydrolysis of gamma-PGA as the main chain and L-PAE as the side chain. In the case of the enzymatic degradation of gamma-PGA nanoparticles with Pronase E, the size of the nanoparticles increased during the initial degradation stage and decreased gradually when the degradation time was extended. Nanoparticles composed of biodegradable amphiphilic gamma-PGA with reactive function groups can undergo further modification and are expected to have a variety of potential pharmaceutical and biomedical applications, such as drug and vaccine carriers.  相似文献   

5.
Krasko MY  Domb AJ 《Biomacromolecules》2005,6(4):1877-1884
The degradation process of poly(ricinoleic-co-sebacic-ester-anhydride)s in buffer solution was investigated by following the composition of the degradation products released into the degradation medium and the degraded polymer. The first week of degradation was characterized by the hydrolysis of the anhydride bonds and significant release of sebacic acid (SA). The remaining oligoesters of SA and ricinoleic acid (RA) degraded into shorter oligoesters composed of RA ester dimers, trimers, and tetramers as well as dimers of RA-SA. To confirm and determine the hydrolytic behavior of the degradation products, short oligoesters of sebacic and ricinoleic acid were synthesized and degraded. It was established that during the hydrolysis under physiological conditions the degradation products have a composition and water absorption similar to those of the synthesized oligoesters.  相似文献   

6.
Reaction of PPN[W(CO)3(R2PC2H4PR2)(SH)] (PPN=Ph3PNPPh3; R=Me, 1; R=Ph, 2) with aromatic aldehydes in the presence of trifluoroacetic acid gave tungsten complexes of thiobenzaldehydes mer-[W(CO)3(R2PC2H4PR2)(η2-SCHR)] (R=Me, 3a-3f; R=Ph, 4a-4e) in high yields. Analogous complexes of aliphatic thioaldehydes mer-[W(CO)3(Me2PC2H4PMe2)(η2-SCHR)] (3g-3l) could only be obtained from the highly electron-rich thiolate complex 1. The structure of 3i (R=i-Bu) was determined by X-ray crystallography. In solution the complexes 3 and 4 are in equilibrium with small quantities of their isomers fac-[W(CO)3(R2PC2H4PR2)(η2-SCHR)]. Reaction of complexes 3 with dimethylsulfate followed by salt metathesis with NH4PF6 gave the alkylation products mer-[W(CO)3(Me2PC2H4PMe2)(η2-MeSCHR)]PF6 (5a-5l) as mixtures of E and Z isomers. The methylated thioformaldehyde complex mer-[W(CO)3(Me2PC2H4PMe2)(η2-MeSCH2)]PF6 (5m) was prepared similarly. Nucleophilic addition of hydride (from LiAlH4) to 5 initially gave thioether complexes mer-[W(CO)3(Me2PC2H4PMe2)(MeSCH2R)] (mer-6) which rapidly isomerized to fac-[W(CO)3(Me2PC2H4PMe2)(MeSCH2R)] (fac-6).  相似文献   

7.
High molar mass random poly(butylene succinate-co-butylene sebacate), P(BS-co-BSe), and poly(butylene succinate-co-butylene adipate), P(BS-co-BA), with different composition, were synthesized and subjected to enzymatic hydrolysis by Lipase from Mucor miehei or from Rhizopus arrhizus. The enzymatic hydrolysis of P(BS-co-BSe)s and P(BS-co-BA)s films produced a mixture of water-soluble monomers and co-oligomers that were separated and identified by on-line high performance liquid chromatography/electrospray ionization mass spectrometry (HPLC/ESI-MS). Optimization of the HPLC analysis allowed the separation of isobar co-oligomers, differing only for the co-monomers sequence. Oligomers with the same monomer composition and molar mass but different sequence were identified by HPLC/ESI-MS-MS on-line analysis. The results obtained show a preferential hydrolytic cleavage induced by the lipases used. In particular, these enzymes prefer cleaving sebacic ester bonds in P(BS-co-BSe) copolymers, whereas succinic ester bonds appear to be hydrolyzed faster than adipic ester bonds in P(BS-co-BA) copolyesters. 1H NMR analysis further substantiates these findings. The primary products generated by lipase hydrolysis of polyester films underwent further degradation at longer reaction times. The HPLC/ESI-MS analysis of these mixtures at various times provided the first evidence that lipase catalysis is active also in water solution, a hydrophobic effect induced by the aliphatic units of these polyesters.  相似文献   

8.
9.
Biodegradable plastics (BPs) have attracted much attention since more than a decade because they can easily be degraded by microorganisms in the environment. The development of aliphatic-aromatic co-polyesters has combined excellent mechanical properties with biodegradability and an ideal replacement for the conventional nondegradable thermoplastics. The microorganisms degrading these polyesters are widely distributed in various environments. Although various aliphatic, aromatic, and aliphatic-aromatic co-polyester-degrading microorganisms and their enzymes have been studied and characterized, there are still many groups of microorganisms and enzymes with varying properties awaiting various applications. In this review, we have reported some new microorganisms and their enzymes which could degrade various aliphatic, aromatic, as well as aliphatic-aromatic co-polyesters like poly(butylene succinate) (PBS), poly(butylene succinate)-co-(butylene adipate) (PBSA), poly(ε-caprolactone) (PCL), poly(ethylene succinate) (PES), poly(l-lactic acid) (PLA), poly(3-hydroxybutyrate) and poly(3-hydoxybutyrate-co-3-hydroxyvalterate) (PHB/PHBV), poly(ethylene terephthalate) (PET), poly(butylene terephthalate) (PBT), poly(butylene adipate-co-terephthalate (PBAT), poly(butylene succinate-co-terephthalate) (PBST), and poly(butylene succinate/terephthalate/isophthalate)-co-(lactate) (PBSTIL). The mechanism of degradation of aliphatic as well as aliphatic-aromatic co-polyesters has also been discussed. The degradation ability of microorganisms against various polyesters might be useful for the treatment and recycling of biodegradable wastes or bioremediation of the polyester-contaminated environments.  相似文献   

10.
Makhatadze and Privalov have analyzed the thermodynamics of transfer of aliphatic and aromatic hydrocarbons from the gas phase into water. Finding that the hydration free energy of aliphatic and aromatic hydrocarbons have different signs, they conclude that the mechanism causing hydrophobicity of these solutes is of a different nature. Here, we offer an alternative analysis of the dissolution of these non-polar compounds into water based on a recently published interpretation scheme for thermodynamic transfer functions. Our analysis shows that the hydrophobicity of aromatic and aliphatic hydrocarbons is qualitatively the same, i.e. its causes are the same namely the extremely high cohesive energy of water which overcomes the favorable solute-solute and solute-water interactions. However, both analyses conclude that the experimentally observed quantitative difference between the interactions of water with aliphatic and aromatic hydrocarbons, can be assigned to the formation of aromatic ring-water H-bonds.  相似文献   

11.
In vitro protein folding of disulfide containing proteins is aided by the addition of a redox buffer, which is composed of a small molecule disulfide and/or a small molecule thiol. In this study, we examined redox buffers containing asymmetric dithiols 1-5, which possess an aromatic and aliphatic thiol, and symmetric dithiols 6 and 7, which possess two aromatic thiols, for their ability to fold reduced lysozyme at pH 7.0 and 8.0. Most in vivo protein folding catalysts are dithiols. When compared to glutathione and glutathione disulfide, the standard redox buffer, dithiols 1-5 improved the protein folding rates but not the yields. However, dithiols 6 and 7, and the corresponding monothiol 8 increased the folding rates 8-17 times and improved the yields 15-42% at 1mg/mL lysozyme. Moreover, aromatic dithiol 6 increased the in vitro folding yield as compared to the corresponding aromatic monothiol 8. Therefore, aromatic dithiols should be useful for protein folding, especially at high protein concentrations.  相似文献   

12.
Gangliotriaosylceramide 3'-sulfate (GgOse3Cer-II3-sulfate) contains the sugar sequence similar to that of GM2 ganglioside except that the NeuAc in GM2 is replaced by a sulfate group. Due to this structural similarity, we have studied the in vitro synthesis of GgOse3Cer-II3-sulfate using the system for GM2. Our results showed that GgOse3Cer-II3-sulfate could be synthesized from lactosylceramide 3'-sulfate and UDP-GalNAc catalyzed by N-acetylgalactosaminyltransferase prepared from rat brain (Dicesare, J. L., and Dain, J. A. (1971) Biochim. Biophys. Acta 231, 385-393). As in the case of GM2, the GgOse3Cer-II3-sulfate biosynthesized in vitro or isolated from rat kidney could also be cleaved by human beta-hexosaminidase A in the presence of GM2-activator (Li, S.-C., Hirabayashi, Y., and Li, Y.-T. (1981) J. Biol. Chem. 256, 6234-6240). The fact that the GM2-activator could stimulate beta-hexosaminidase A to hydrolyze both GM2 and Gg-Ose3Cer-II3-sulfate indicates that these two glycolipids may be catabolyzed by the same mechanism.  相似文献   

13.
14.
15.
16.
Dimethoate-degrading enzymatic activity in Bacillus licheniformis, Pseudomonas aeruginosa, Aeromonas hydrophila, Proteus mirabilis and Bacillus pumilus was found to be 6.4, 1.760, 4.09, 1.196 and 0.505 units/mg protein, respectively. The Escherichia coli C600 transconjugants of the isolated bacterial strains also exhibited dimethoate-degrading enzymatic activities. The cured derivatives did not show any decrease in the amount of dimethoate substrate and did not harbour plasmid as found in the original and transconjugant strains. Thus, the ability of enzymatic degradation of dimethoate was plasmid-mediated in B. licheniformis, Ps. aeruginosa, A. hydrophila, P. mirabilis and B. pumilus.  相似文献   

17.
The gene encoding an aromatic and neutral aliphatic amino acid permease of Penicillium chrysogenum was cloned, functionally expressed and characterized in Saccharomyces cerevisiae M4276. The permease, designated PcMtr, is structurally and functionally homologous to Mtr of Neurospora crassa, and unrelated to the Amino Acid Permease (AAP) family which includes most amino acid permeases in fungi. Database searches of completed fungal genome sequences reveal that Mtr type permeases are not widely distributed among fungi, suggesting a specialized function.  相似文献   

18.
Organodiselenides are potential building blocks for a new class of self-assembled monolayers e.g. on coinage metals. For this, the diselenides have to be extremely pure and in particular must not contain any traces of higher organoselenides. Several strategies for the preparation of organodiselenides were compared in their efficiency in particular regarding the purity of the products. It could be shown that most literature-established procedures generate significant amounts of triselenides which are hard to detect with conventional analytic methods. While in case of the aromatic diselenides the direct synthesis via the corresponding Grignard reagent resulted in pure diselenides, the preparation of the aliphatic and araliphatic derivatives required a more indirect route through the corresponding selenocyanates.  相似文献   

19.
20.
The bacterial degradation of catechol, 3-methylcatechol, 2,3-dihydroxy-β-phenylpropionic acid, and protocatechuic acid has been studied in detail. From the results obtained a general sequence has been proposed for the microbial oxidation of dihydroxy aromatic compounds.  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