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.
Deng C  Tian H  Zhang P  Sun J  Chen X  Jing X 《Biomacromolecules》2006,7(2):590-596
Advances in tissue engineering require biofunctional scaffolds that can provide not only physical support for cells but also chemical and biological cues needed in forming functional tissues. To achieve this goal, a novel RGD peptide grafted poly(ethylene glycol)-b-poly(L-lactide)-b-poly(L-glutamic acid) (PEG-PLA-PGL/RGD) was synthesized in four steps (1) to prepare diblock copolymer PEG-PLA-OH and to convert its -OH end group into -NH(2) (to obtain PEG-PLA-NH(2)), (2) to prepare triblock copolymer PEG-PLA-PBGL by ring-opening polymerization of NCA (N-carboxyanhydride) derived from benzyl glutamate with diblock copolymer PEG-PLA-NH(2) as macroinitiator, (3) to remove the protective benzyl groups by catalytic hydrogenation of PEG-PLA-PBGL to obtain PEG-PLA-PGL, and (4) to react RGD (arginine-glycine-(aspartic amide)) with the carboxyl groups of the PEG-PLA-PGL. The structures of PEG-PLA-PGL/RGD and its precursors were confirmed by (1)H NMR, FT-IR, amino acid analysis, and XPS analysis. Addition of 5 wt % PEG-PLA-PGL/RGD into a PLGA matrix significantly improved the surface wettability of the blend films and the adhesion and proliferation behavior of human chondrocytes and 3T3 cells on the blend films. Therefore, the novel RGD-grafted triblock copolymer is expected to find application in cell or tissue engineering.  相似文献   

2.
We report on the modulation of phase morphology, plasticization properties, and thermal stability of films of partly branched poly(l-lactide)-co-poly(epsilon-caprolactone) copolymer (PLLA-co-PCL) with additions of low molecular weight compounds, namely, triethyl citrate ester, diethyl phthalate, diepoxy polyether (poly(propylene glycol) diglycidyl ether), and with epoxidized soybean oil (ESO). The PLLA-co-PCL/polyether films showed significant stability against thermal depolymerization, high film flexibility, and good plasticizing properties, probably due to cross-linking and chain branching formation between diepoxy groups with both the end carboxyl and hydroxyl groups of the PLLA copolymer (initially present or generated during the degradation process) to produce primary ester and ether bonds, respectively. Diethyl phthalate and triethyl citrate ester were found to be efficient plasticizers for PLLA copolymer in terms of glass transition and mechanical properties, but the more water-soluble plasticizer triethyl citrate induced a dramatic loss in the molecular weight of the copolymer. Although ESO cannot play the role of a plasticizer, it substantially stabilizes and retards thermal depolymerization of the PLLA copolymer matrix, possibly because of a reaction between epoxy groups with the end carboxyl and hydroxyl groups of the PLLA copolymer. The presence of ESO in PLLA-co-PCL/ESO/triethyl citrate blends enhanced the compatibility and miscibility of the plasticizer with the PLLA copolymer matrix, considerably improved the mechanical properties (elongation at break), and substantially stabilized the copolymer against thermal depolymerization. It seems likely that the epoxy groups interact not only with the end hydroxyl and carboxyl group of the copolymer but as well with the hydroxyl group of triethyl citrate plasticizer to produce a new ether bond (C-O-C) as the cross-linking unit. On the other hand, for PLLA-co-PCL/ESO/polyether blends, (80/10/10) epoxidized oil distorts the compactness of the blend by diminishing the proposed entanglements between carboxyl, hydroxyl, and diepoxy groups of polyether and reduces the high elongation properties otherwise observed in the PLLA-co-PCL/polyether films. The multicomponent approach toward modulating poly(l-lactide)-co-poly(epsilon-caprolactone) copolymer films using epoxy compounds and plasticizers and the insight into the nature of various PLLA matrixes presented here offer advantages to a broad engineering of PLLA copolymer films having desirable physical properties and multiphase behavior for efficient uses in future technical applications.  相似文献   

3.
The use of biodegradable derivatives of poly(organo phosphazenes) for the preparation of nanoparticles and their surface modification with the novel poly(ethylene oxide) derivative of poly(organo phosphazene) has been assessed using a range of in vitro characterization methods. The nanoparticles were produced by the precipitation solvent evaporation method from the derivative co-substituted with phenylalanine and glycine ethyl ester side groups. A reduction in particle size to less than 200 nm was achieved by an increase in pH of the preparation medium. The formation (and colloidal stability) of these nanoparticles seems to be controlled by two opposite effects: attractive hydrophobic interactions between phenylalanine ester groups and electrostatic repulsions arising from the carboxyl groups formed due to (partial) hydrolysis of the ester bond(s) at the high pH of the preparation medium. The poly[(glycine ethyl ester)phosphazene] derivative containing 5000-Da poly(ethylene oxide) as 5% of the side groups was used for the surface modification of nanoparticles. Adsorbed onto the particles, the polymer produced a thick coating layer of approximately 35 nm. The coated nanoparticles exhibited reduced surface negative potential and improved colloidal stability toward electrolyte-induced flocculation, relative to the uncoated system. However, the steric stabilization provided was less effective than that of a Poloxamine 908 coating. This difference in effectiveness of the steric stabilization might indicate that, although both the stabilizing polymers possess a 5000-Da poly(ethylene oxide) moiety, there is a difference in the arrangements of these poly(ethylene oxide) chains at the particle surface. (c) 1996 John Wiley & Sons, Inc.  相似文献   

4.
J C Wu  J Stubbe  J W Kozarich 《Biochemistry》1985,24(26):7569-7573
Incubation of poly(dA-[3'-3H]dU), poly(dA-[5'-3H]dU), or poly(dA-[5'-3H]dT) under a variety of conditions with activated bleomycin resulted in the production of free nucleic acid base, base propenal, and a small amount of 3H2O. Adjustment of the terminated reaction mixture to pH 10 and incubation at 95 degrees C resulted in a time-dependent increase in 3H2O to an amount equal to the amount of free base. If the terminated reaction mixture was incubated with NaBH4 prior to the heat and alkaline treatment, the release of 3H2O was significantly inhibited. These results are consistent with the generation by activated bleomycin of a 4'-ketone yielding free base, with the exchange of the 3'- and 5'-hydrogens by enolization and with the alkaline-induced strand scission occurring from this intermediate.  相似文献   

5.
Solution conformation of poly(L-lysyl-L-glutamic acid) (PLGU) and poly(L-lysyl-L-glutamine) (PLGN) was studied in water as a function of pH, added salt, detergents, methanol and trifluoroethanol (TFE). Both the polypeptides exhibit no ordered conformation in the pH range 1.5-12.5; salts and detergents did not have any marked effect. Replacement of side chain carboxyl by an amide group did not help in inducing PLGN to adopt a helical conformation even at pH as high as 12.0, unlike poly(L-lysine). The helicogenic solvents, methanol and TFE, induce formation of weak helices in PLGU as well as in PLGN. It is not unlikely that H-bonding between the side chains leads to stabilizing an unordered conformations.  相似文献   

6.
After incubating root apices from two-day-old bean seedlings with [3H] adenine the RNA was extracted from whole cells or polysomes, and the poly (A) sequences were isolated by nuclease digestion followed by poly(U)-Sepharose chromatography. The alterations of the RNA molecules due to the various treatments were monitored by sucrose density gradients. It was found that sequential extraction first at pH 7.6 then at pH 9.0 did not result in a separation between RNA poor in poly(A) sequences and poly(A)-rich RNA. Furthermore chromatography analysis of hydrolysates from nuclease-resistant RNA extracted either at pH 7.6 or pH 9.0 revealed that AMP constituted nearly 95% of the bases and that the poly(A) sequences, about 200 bases, were located at the 3' terminus of the polyadenylated RNA. No size difference was found for the poly(A) segment between the pH-7.6-extracted RNA and that extracted at pH 9.0.  相似文献   

7.
Li J  Wang T  Wu D  Zhang X  Yan J  Du S  Guo Y  Wang J  Zhang A 《Biomacromolecules》2008,9(10):2670-2676
Synthesis of novel zwitterionic block copolypeptides, poly(N-isopropylacrylamide)-block-poly(L-glutamic acid-co-L-lysine) [PNiPAM(n)(PLG(x)-co-PLLys(y))m , where n is the number-average degree of polymerization (DP(n)) of PNiPAM block, x and y are the mole fraction of glutamic acid and lysine residues, respectively, and m is the total DP(n) of the peptide block], and their stimuli-responsiveness to temperature and pH variation in aqueous solutions are described. Initiated with the amino-terminated poly(N-isopropylacrylamide) (PNiPAM(n)-NH2), ring-opening polymerization (ROP) of a mixture of gamma-benzyl-L-glutamate N-carboxyanhydride (BLG-NCA), and Boc-L-lysine N-carboxyanhydride (BLLys-NCA) afforded the block copolypeptides PNiPAM(n)(PBLG(x)-co-PBLLys(y))m, with a poly(N-isopropylacrylamide) block together with a random copolypeptide block, which was then deprotected with HBr/trifluoroacetic acid into the double hydrophilic block copolypeptides, PNiPAM(n)(PLG(x)-co-PLLys(y))m. Their block ratios and lengths, as well as the amino acid residue ratios in the random copolypeptide block are varied (n = 360, x = 0.4-0.5, y = 0.4-0.6, and m = 220-252). The secondary structures of the copolypeptides in aqueous solution at different pH conditions were examined. Phase transitions in aqueous solutions induced by both pH and temperature variation were investigated by (1)H NMR spectroscopy. The transitions induced by temperature were also explored by turbidity measurements using UV/vis spectroscopy for their lower critical aggregation temperature (LCAT) determination. Furthermore, these aggregation processes were followed by dynamic light scattering measurements.  相似文献   

8.
In this work, poly(aniline-co-o-aminophenol) (copolymer) was used as an electron transfer mediator in the electrochemical oxidation of catechol due to its reversible redox over a wide range of pH. The experimental results indicate that the anodic peak potential of catechol at the copolymer electrode is lower than that at the platinum electrode in a solution consisting of catechol and sodium sulfate with pH 5.0, and the activation energy for the electrochemical oxidation of catechol at the copolymer electrode is low (23.6 kJ mol(-1)). These are strong evidence for the electrocatalytic oxidation of catechol at the copolymer electrode. The -OH group on the copolymer chain plays an important role in the electron transfer between the copolymer electrode and catechol in the solution. Based on the catalytic oxidation, the copolymer is used as a sensor to determine the concentration of catechol. The response current of the sensor depends on the concentration of catechol, pH, applied potential and temperature. At 0.55 V (versus saturated calomel reference electrode (SCE)) and pH 5.0, the sensor has a fast response (about 10s) to catechol and good operational stability. The sensor shows a linear response range between 5 and 80 microM catechol with a correlation coefficient of 0.997. It was found that phenol and resorcinol cannot be oxidized at the copolymer electrode at potentials < or =0.55 V, so controlling the sensor potential affords a good way of avoiding the effect of phenol and resorcinol on the determination of catechol.  相似文献   

9.
The uptake and metabolism of H-Pro-[3H]Leu-Gly-NH2 ([3H]PLG) in rat brain was investigated by reverse-phase paired-ion high pressure liquid chromatography. Following in vitro incubation of [3H]PLG with rat brain subcellular preparations, the microsomal-cytosol fraction was about twice as active in degrading PLG as the crude mitochondrial-synaptosomal fraction. For both enzyme preparations the pH optimum was found at pH 7-7.5. The major labeled metabolite was [3H]leucine, whereas 3H]labeled Leu-Gly-NH2 as the only labeled peptide intermediate was found in trace amounts. After intravenous injection of [3H]PLG the uptake of unmetabolized peptide in the brain appeared to be very low: 0.008% and 0.001% of the administered dose/g tissue at 2 and 5 min after injection respectively, while at longer survival times intact peptide was below the detection limit. Compared with the intravenous route of administration, intracerebroventricular injection of [3H]PLG yielded much higher brain concentrations of unmetabolized PLG. Following both routes of administration, the metabolite profile was in agreement with that obtained after in vitro incubation. However, the in vivo experiments also showed considerable incorporation of [3H]leucine liberated from [3H]PLG into proteins. Both the in vitro and in vivo results indicate that the initial cleavage of PLG in rat brain occurs at the NH2-terminus and that the dipeptide intermediate H-Leu-Gly-NH2 is subsequently hydrolyzed to its constituent amino acids very rapidly.  相似文献   

10.
A T-jump investigation of the binding of Cyan40 [3-methyl-2-(1,2,6-trimethyl-4(1H)pyridinylidenmethyl)-benzothiazolium ion] and CCyan2 [3-methyl-2-[2-methyl-3-(3-methyl-2(3H)-benzothiazolylidene)-1-propenyl]-benzothiazolium ion] with poly(dA-dT) x poly(dA-dT) and poly(dG-dC) x poly(dG-dC) is performed at I = 0.1M (NaCl), 25 degrees C and pH 7. Two kinetic effects are observed for both systems. The binding process is discussed in terms of the sequence D + P <==> P,D <==> PD(I) <==> PD(II), which leads first to fast formation of a precursor complex P,D and then to a partially intercalated complex PD(I) which converts to the fully intercalate complex PD(II). Concerning CCyan2 the rate parameters depend on the polymer nature and their analysis shows that in the case of poly(dG-dC) x poly(dG-dC) the most stable bound form is the fully intercalated complex PD(II), whereas in the case of poly(dA-dT) x poly(dA-dT) the partially intercalated complex PD(I) is the most stable species. Concerning Cyan40, the rate parameters remain unchanged on going from A-T to G-C indicating that this dye is unselective.  相似文献   

11.
The graft copolymer, poly(maleic anhydride/styrene)-co-polyethylene was prepared. The copolymer immobilized bovine serum albumin (BSA), but the amount coupled appeared to be effected by the amount of styrene in the graft copolymer, temperature, and pH of the coupling medium. Competition existed between hydrolysis of the grafted anhydride groups and the protein. A graft copolymer with 66% add-on immobilized 4.5 mg/glucose oxidase/g copolymer, 4.6 mg alkaline phosphates/g copolymer and 0.2 mg cell of Bacillus stearothermophilus/g copolymer. A number of copolymers containing poly(maleic anhydride/vinyl acetate)-co-polyethylene were prepared to cover a range of grafting levels. These immobilized larger quantities of BSA, alkaline phosphatase, and cells of B. stearothermophilus than did the styrene graft copolymer. The copolymer was also hydrolyzed to release the hydroxyl group from the poly(vinyl acetate) component of the grafted chains. Using p-benzoquinone as the "activating agent," the copolymer coupled to BSA and to acid phosphatase. Using p-toluene-sulfonyl chloride, the copolymer was very effective in immobilizing trypsin.  相似文献   

12.
Addition of poly(A) to nuclear RNA occurs soon after RNA synthesis   总被引:11,自引:2,他引:9       下载免费PDF全文
A kinetic analysis of the appearance of [3H]uridine label in RNA sequences that neighbor poly(A), as well as the incorporation of [3H]adenosine label into both the RNA chain and the poly(A) of poly(A)-containing molecules, shows that poly(A) is added within a minute or so after RNA chain synthesis in Chinese hamster ovary cells and HeLa cells. Previous conclusions by several groups (5-7) that poly(A) might be added as long as 20-30 min after RNA synthesis appear to be in error, and the present conclusion seems much more in line with several different types of recent studies with specific mRNAs that suggest prompt poly(A) addition (13-16).  相似文献   

13.
The in vitro hydrolytic degradation of hydroxyl-functionalized poly(alpha-hydroxy acid)s was investigated. Benzyl-ether-protected hydroxyl-functionalized dilactones (S)-3-benzyloxymethyl-(S)-6-methyl-1,4-dioxane-2,5-dione (1a) and (S)-3-benzyloxymethyl-1,4-dioxane-2,5-dione (1b) were copolymerized in a melt with various amounts of L-lactide using benzyl alcohol and SnOct2 as the initiator and catalyst, respectively. The benzyl groups were removed by hydrogenation to yield polyesters with hydroxyl functional groups, poly(lactic acid-co-hydroxymethyl glycolic acid) and poly(lactic acid-co-glycolic acid-co-hydroxymethyl glycolic acid) (2a and 2b). Degradation of the hydroxyl-functionalized polyesters and poly(lactic-co-glycolic acid) (50/50) was studied by incubation of pellets of these polymers in phosphate buffer (174 mM, pH 7.4) at 37 degrees C. Polymer degradation was monitored by mass-loss measurements and by gel permeation chromatography, differential scanning calorimetry, and 1H NMR analysis. The degradation times ranging from less than 1 day (for the homopolymer of 2a) to 2 months (copolymer of 25% 2a and 75% lactide) were found. The degradation rates increased with increasing hydroxyl density of the polymers, which was associated with a switch from bulk to surface erosion. NMR and thermal analysis showed that the moieties with the hydroxyl groups were preferentially removed from the degrading polymer. In conclusion, this study shows that the degradation rate of polyesters containing 2a and 2b can be tailored from a few days to 2 months, making them very suitable for biomedical and pharmaceutical applications.  相似文献   

14.
Positively charged molecules such as protamine, leukocyte cationic protein, and the carboxyl terminus of platelet factor 4 have been shown to increase fibrin fiber thickness. Synthetic homo poly(L-amino acids) were used to explore the role of charge and molecular weight of cationic molecules on fibrin assembly. The effects of poly(L-lysine) (PLL), poly(L-glutamic acid) (PLG), poly(L-aspartic acid) (PLA), poly(L-histidine) (PLH), and poly(L-arginine) (PLArg) on the assembly and structure of fibrin gels were studied by using light-scattering techniques. At a PLG (Mr 60,000) concentration of 80 micrograms/mL and a PLA (Mr 20,000) concentration of 64 microgram/mL, neither of these negatively charged polymers produced a detectable change in either fibrin assembly kinetics or final structure. Positively charged PLArg (16 micrograms/mL) caused a 30% increase in fibrin fiber mass/length ratio without calcium. In contrast, PLH (16 micrograms/mL), also positively charged, had no effect in the absence of CaCl2 but produced a 40% increase in fiber mass/length ratio with 5 mM CaCl2. At concentrations as low as 1 microgram/mL, positively charged PLL increased the initial fibrin assembly kinetics and led to larger fiber mass/length ratio. The impact on fibrin mass/length ratio was equivalent for three different molecular weight preparations of PLL (Mr 25,000, 90,000, and 240,000). The lack of a molecular weight effect on fiber thickness and the low polymer concentrations required to produce the perturbation argue against an excluded volume effect as the mechanism by which lateral fiber growth is augmented. Mechanisms by which poly(L-amino acids) may perturb fibrin assembly are discussed.  相似文献   

15.
Novel synthetic collagen fibers, poly(PHG) made by polycondensation of Pro-Hyp-Gly, spontaneously assume polymeric structure with molecular weights greater than 105. Its application for biomaterials has been explored, but that for a platelet agonist has not been investigated. Poly(PHG)-induced platelet aggregation independently of thromboxane A2 and integrin α2β1. Poly(PHG)-induced tyrosine phosphorylation of glycoprotein VI (GPVI)-related molecules and failed to activate GPVI/FcRγ-deficient platelets. Binding of GPVI to poly(PHG) was confirmed by a surface plasmon resonance spectroscopy, suggesting that poly(PHG) activates platelets through GPVI. Poly(PHG) is an useful research tool to investigate GPVI-mediated signals and a substitute for collagen in platelet functional assays.  相似文献   

16.
A S Acharya  R Seetharam 《Biochemistry》1985,24(18):4885-4890
X-ray diffraction analysis of deoxyhemoglobin S crystals has implicated that a number of carboxyl groups of the protein are present at or near the intermolecular contact regions. The reactivity of these or other carboxyl groups of hemoglobin S for the amidation with an amino sugar, i.e., glucosamine, and the influence of amidation on the oxygen affinity and polymerization have been investigated. Reaction of oxyhemoglobin S at pH 6.0 and 23 degrees C with 20 mM 1-ethyl-3-[3-(dimethylamino)propyl]carbodiimide (EDC) and 100 mM [3H]glucosamine for 1 h resulted in an incorporation of nearly two residues of glucosamine per tetramer. The amidation was very specific for the carboxyl groups of globin; the glucosamine was not incorporated into the heme carboxyls. Derivatization of hemoglobin S by glucosamine increased the O2 affinity of the protein but had no influence on either the Hill coefficient or the Bohr effect. Amidation by glucosamine also increased the solubility of deoxyhemoglobin S by about 55%. Tryptic peptide mapping of the modified hemoglobin S indicated that the peptides beta-T3 and beta-T5 contained the glucosamine incorporated into the protein. Sequence analysis of glucosamine-modified beta-T3 and beta-T5 demonstrated that the gamma-carboxyl groups of Glu-22 and Glu-43, respectively, had been derivatized with glucosamine. The residue Glu-43(beta) shows a high selectivity toward glycine ethyl ester also, whereas Glu-22(beta) is not reactive toward this amine.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
1. The rate of reaction of mustard gas (H) with thirteen proteins has been determined. The extreme variation in reaction rates is about 100:1. 2. No qualitative difference in the results was observed when the treatment with H was carried out by the Dixon or stirring methods. 3. The kinetics have been analyzed and a bimolecular equation derived which fits the facts. 4. The carboxyl groups of all proteins reacted when the reaction with H was carried out at pH 6.0 in M/25 acetate buffer. In most cases the number of carboxyl groups covered was approximately equal to the number of H residues bound. 5. The amino groups of proteins failed to react with the possible exception of yeast hexokinase. 6. The color obtained when proteins were mixed with Folin''s phenol reagent at pH 8.0 decreased as the protein was treated with H. The color returned on treatment of the H-protein with alkali and many of the combined H groups were hydrolyzed. Similar results were observed when a concentrated glycyltyrosine solution was treated with H.  相似文献   

18.
The binding parameters (K, omega) and the free energy (DeltaG(0)) of triple helix formation have been estimated for complexes of oligo(U)(n) (n = 5, 7-10) with poly(A) . poly(U) on the basis of hypochromicity measurements. The data were treated according to the formula of McGhee and von Hippel [J. Mol. Biol. 86 (1974) 469] by a computer program ALAU [H. Schütz et al., Stud. Biophys. 104 (1984) 23] which takes absorbancies and total concentrations as input. In 1 mM cacodylate buffer pH 7.0 with 10 mM NaCl and 10 mM MgCl(2) at 5 degrees C the free energy of contiguous binding was found to be a linear function of the oligomer length with a slope of DeltaG(c,U)(0) = -0.72 (+/-0.03) kcal x mol(-1) per nucleotide. The mean cooperativity coefficient (omega) was 24.5 (+/- 5.6), and the corresponding free energy of interaction between the neighbouring oligonucleotides in the third strand was DeltaG(0(omega)) = -1.74 (+/-0.13) kcal x mol(-1).  相似文献   

19.
Zeng F  Lee H  Chidiac M  Allen C 《Biomacromolecules》2005,6(4):2140-2149
Novel amphiphilic six-arm star diblock copolymers based on biocompatible and biodegradable poly(delta-valerolactone) (PVL) and methoxy poly(ethylene glycol) (MePEG) were synthesized by a two-step process. First, the hydrophobic star-shaped PVL with hydroxyl terminated functional groups was synthesized using a multifunctional alcohol, dipentaerythritol (DPE), as the initiator and fumaric acid as the catalyst. The amphiphilic six-arm star copolymer of poly(delta-valerolactone)-b-methoxy poly(ethylene glycol), (PVL-b-MePEG)(6), was then synthesized by coupling the hydroxyl terminated six-arm PVL homopolymer with alpha-methoxy-omega-chloroformate-poly(ethylene glycol) (MePEG-COCl). (1)H NMR and GPC analyses confirmed the successful synthesis of star-shaped copolymers with predicted compositions and narrow molecular weight distributions. DSC analysis revealed that the glass transition temperatures of the star PVL homopolymers with M(n) between 5000 and 49 000 are not dependent on their molecular weights, whereas the melting temperatures of both the PVL homopolymers and the amphiphilic (PVL-b-MePEG)(6) copolymers increase with an increase in the PVL molecular weight. Micelles were prepared from the (PVL-b-MePEG)(6) copolymers via the dialysis method and found to have effective mean diameters ranging from 10 to 45 nm, depending on the copolymer composition. In addition, the (PVL-b-MePEG)(6) copolymers having lower PVL content were found to form micelles with a narrow monomodal size distribution, whereas the copolymers having higher PVL content tended to form aggregates with a bimodal size distribution. The noncytotoxicity of the copolymers was also confirmed in CHO-K1 fibroblast cells using a cell viability assay, indicating that the (PVL-b-MePEG)(6) copolymers are suitable for biomedical applications such as drug delivery.  相似文献   

20.
The reaction mechanism of automodification of poly (ADP-ribose) synthetase was studied. The synthetase, bound to nicked DNA-cellulose in a small column, was pulse-labelled with [3H]NAD in the presence of Mg2+, and then chased with [14C]NAD under the same conditions after complete washing of [3H]NAD. The poly(ADP-ribose), synthesized on the synthetase molecule, was digested with snake venom phosphodiesterase and analyzed. The [3H]-labeled product (35% of the total product) was identified as isoADP-ribose but [3H]-labelled AMP was not detected. The average chain length was 16.0 and the terminal AMP was detected as [14C]-labelled AMP. These results indicate that the initially attached ADP-ribose unit at an automodification site was successively elongated by the addition of a new ADP-ribose unit to the terminal AMP moiety.  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