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.
The synthesis and characterization of water-soluble random copolymers containing L -asparagine with N5-(4-hydroxybutyl)-L -glutamine, and the thermally induced helix-coil transitions of these copolymers in water, are described. The incorporation of L -asparagine was found to decrease the helix content of the polymers in water at all temperatures. The Zimm-Bragg parameters σ and s for the helix-coil transition in poly(L -asparagine) in water were deduced from an analysis of the copolymer melting curves in the manner described in earlier papers. The computed values of s indicate that asparagine destabilizes helical sequences at all temperatures in the range 0–60°C.  相似文献   

2.
Water-soluble, random copolymers containing L -glutamine and either N5-(3-hydroxypropyl)-L -glutamine or N5-(4-hydroxybutyl)-L -glutamine were synthesized, fractionated, and characterized. The thermally induced helix–coil transitions of these copolymers were studied in water. A short-range interaction theory was used to deduce the Zimm-Bragg parameters σ and s for the helix–coil transition in poly(L -glutamine) in water from an analysis of the melting curves of the copolymers in the manner described in earlier papers. The computed values of s indicate that L -glutamine is helix-indifferent at low temperature and a helix-destabilizing residue at high temperature in water. At all temperatures in the range of 0–70°C, the glutamine residue promotes helix–coil boundaries since the computed value of σ is large.  相似文献   

3.
A Zagari  G Némethy  H A Scheraga 《Biopolymers》1990,30(9-10):961-966
The alteration of polymer conformational properties caused by the replacement of L-proline by L-azetidine-2-carboxylic acid (Aze) has been studied by means of conformational energy computations. In addition to poly(Aze), two sequential copolymers, poly(Pro-Aze) and poly(Aze3-Pro3), have been investigated. All polymers containing Aze are more flexible than poly(Pro). This is a consequence of an increased number of permitted conformational states for the Aze residue, as compared to Pro, when they are incorporated into a polypeptide, as well as of a lessened cooperativity of the trans-cis transition. The results of the computation can be used to interpret the observed physical properties of poly(Aze) and of its copolymers.  相似文献   

4.
During the benzyloxycarbonylation of H-Asp(OBzl)-OH by the Schotten-Bauman reaction with benzyloxycarbonyl chloride in the presence of NaHCO3 or Na2CO3, besides Z-Asp(OBzl)-OH, Z-Asp(OBzl)-Asp(OBzl)-OH was formed as side product, although the extent of the dipeptide formation differed depending on the base used (10% and 20% respectively). It was found that melting point, rotation value and Rf values upon thin-layer chromatography of Z-Asp(OBzl)-Asp(OBzl)-OH were quite similar to those of Z-Asp(OBzl)-OH.  相似文献   

5.
Previously used procedures for processing the amino acids from 6N hydrochloric acid hydrolysis of poly[N5-(4-hydroxybutyl)-L -glutamine], poly[N5-(3-hydroxypropyl)-L -glutamine], and several random copolymers derived from these, led to the formation of spurious products. These have been isolated and characterized as the γ-ester of glutamic acid and the hydroxyalkyl amine, and chloro-alkyl amine hydrochloride. The former reduces the observed values for glutamic acid, but the latter has no effect on them. A method is used to avoid formation of these artifacts in the amino-acid analysis. Of all the copolymers studied previously in this series, the compositions of only those containing L -serine are in error as a result of the formation of the γ-ester. A redetermination of the amino-acid compositions of the copolymer fractions studied earlier leads to slightly revised values for the Zimm-Bragg parameters σ and s of serine.  相似文献   

6.
The solid-state conformation of copolymers of β-benzyl-L -aspartate [L -Asp(OBzl)] with L -leucine (L -Leu), L -alanine (L -Ala), L -valine (L -Val), γ-benzyl-L -glutamate [L -Glu(OBzl)], or ?-carbobenzoxy-L -lysine (Cbz-L -Lys) has been studied by ir spectroscopy and circular dichroism (CD). The ir spectra in the region of the amide I and II bands and in the region of 700–250 cm?1 have been determined. The results from the ir studies are in good agreement with data obtained by CD experiments. Incorporation of the amino acid residues mentioned above into poly[L -Asp(OBzl)] induces a change from the left-handed into the right-handed α-helix. This conformational change for the poly[L -Asp(OBzl)] copolymers was observed in the following composition ranges: L -Leu, 0–15 mol %; L -Ala, 0–32 mol %; L -Val, 0–8 mol %; L -Glu(OBzl), 3–10 mol %; and Cbz-L -Lys, 0–9 mol %.  相似文献   

7.
R K Dukor  T A Keiderling 《Biopolymers》1991,31(14):1747-1761
The "random coil" conformational problem is examined by comparison of vibrational CD (VCD) spectra of various polypeptide model systems with that of proline oligomers [(Pro)n] and poly(L-proline). VCD, ir and uv CD spectra of blocked L-proline oligopeptides [(Pro)n, n = 2-12] in different solvents are reported and compared to the spectra of poly(L-proline) II, poly(L-glutamic acid), and unblocked proline oligomers. Based on the chain-length dependence of the VCD and electronic CD (ECD) spectra of proline oligomers, it is established that VCD spectra are dominated by short-range interactions. The VCD of random coil model polypeptides is shown to be identical in shape but smaller in magnitude than poly(L-proline) II and of similar magnitude to that of (Pro)n (n = 3, 4). Based on the spectral evidence, it is concluded that the "random coil" conformation has a large fraction of helical regions, conformationally similar to the left-handed, 3(1) polyproline II helix, as was previously suggested by Krimm and co-workers. This conclusion is further supported by studies of effects of salt (CaCl2, LiBr, LiClO4), temperature (5-75 degrees C), and pH on the VCD spectra of L-proline oligomers, poly(L-proline) II, and poly(L-glutamic acid). These show that, after each of these perturbations, a significant local ordering remains in the oligomers and polymers studied, and that charged polypeptides such as poly(L-glutamic acid) are more flexible than are polyproline or even L-proline oligomers.  相似文献   

8.
Synthesis is described of four tuftsin derivatives containing a D-glucopyranosyl or a D-galactopyranosyl unit covalently linked to the hydroxy side chain function of the threonine residue through either an alpha or beta O-glycosidic linkage. Fmoc-threonine derivatives containing the suitable unprotected sugar were used for incorporating the O-glycosylated amino acid residue. Z-Thr[alpha-Glc(OBzl)4]-OBzl and Z-Thr[alpha-Gal(OBzl)4]-OBzl were prepared from the tetra-O-benzylated sugar and Z-Thr-OBzl by the trichloroacetimidate method in the presence of trimethylsilyl trifluoromethane sulfonate. The alpha glycosylated threonine derivatives were converted into Fmoc-Thr(alpha-Glc)-OH and Fmoc-Thr(alpha-Gal)-OH by catalytic hydrogenation followed by acylation with Fmoc-OSu. beta-Glucosylation and beta-galactosylation of threonine were carried out by reacting the proper per-O-acetylated sugar with Z-Thr-OBzl and boron trifluoride ethyl etherate in dichloromethane. Catalytic hydrogenation of the beta-O-glycosylated threonine derivatives followed by acylation with Fmoc-OSu and deacetylation with methanolic hydrazine yielded Fmoc-Thr(beta-Glc)-OH and Fmoc-Thr(beta-Gal)-OH, respectively. The O-glycosylated threonine derivatives were then reacted with H-Lys(Z)-Pro-Arg(NO2)-OBzl in the presence of DCC and HOBt and the resulting glycosylated tuftsin derivatives were fully deblocked by catalytic hydrogenation, purified by HPLC, and characterized by optical rotation, amino acid analysis, and 1H NMR. The beta-galactosylated tuftsin was also prepared by the continuous flow solid phase procedure.  相似文献   

9.
Prolyl 4-hydroxylase (EC 1.14.11.2) catalyzes the hydroxylation of -X-Pro-Gly- sequences and plays a central role in the synthesis of all collagens. The [alpha(I)]2beta2 type I enzyme is effectively inhibited by poly(L-proline), whereas the [alpha(II)]2beta2 type II enzyme is not. We report here that the poly(L-proline) and (Pro-Pro-Gly)10 peptide substrate-binding domain of prolyl 4-hydroxylase is distinct from the catalytic domain and consists of approximately 100 amino acids. Peptides of 10-19 kDa beginning around residue 140 in the 517 residue alpha(I) subunit remained bound to poly(L-proline) agarose after limited proteolysis of the human type I enzyme tetramer. A recombinant polypeptide corresponding to the alpha(I) subunit residues 138-244 and expressed in Escherichia coli was soluble, became effectively bound to poly(L-proline) agarose and could be eluted with (Pro-Pro-Gly)10. This polypeptide is distinct from the SH3 and WW domains, and from profilin, and thus represents a new type of proline-rich peptide-binding module. Studies with enzyme tetramers containing mutated alpha subunits demonstrated that the presence of a glutamate and a glutamine in the alpha(II) subunit in the positions corresponding to Ile182 and Tyr233 in the alpha(I) subunit explains most of the lack of poly(L-proline) binding of the type II prolyl 4-hydroxylase. Keywords: collagen/dioxygenases/peptide-binding domain/ proline-rich/prolyl hydroxylase  相似文献   

10.
A number of taxonomically-related bacteria have been identified which accumulate poly(hydroxyalkanoate) (PHA) copolymers containing primarily 3-hydroxyvalerate (3HV) monomer units from a range of unrelated single carbon sources. One of these, Rhodococcus sp. NCIMB 40126, was further investigated and shown to produce a copolymer containing 75 mol% 3HV and 25 mol% 3-hydroxybutyrate (3HB) from glucose as sole carbon source. Polyesters containing both 3HV and 3HB monomer units, together with 4-hydroxybutyrate (4HB), 5-hydroxyvalerate (5HV) or 3-hydroxyhexanoate (3HHx), were also produced by this organism from certain accumulation substrates. With valeric acid as substrate, almost pure (99 mol% 3HV) poly(3-hydroxyvalerate) was produced. N.m.r. analysis confirmed the composition of these polyesters. The thermal properties and molecular weight of the copolymer produced from glucose were comparable to those of PHB produced by Alcaligenes eutrophus.  相似文献   

11.
Poly(methyl methacrylate) and polystyrene having terminal amino groups were synthesized by the radical polymerization of those monomers in the presence of 2-mercaptoethylammonium chloride as a chain-transfer agent. By the terminal group analysis and the molecular weight determination of the polymers, 0.5–1.3 amino groups were found in a chain of poly(methyl methacrylate) and 0.5–2.5 amino groups in a chain of polystyrene. Using these polymers having a terminal amino group as an initiator, the block polymerization of α-amino acid N-carboxyanhydride (NCA) was carried out. In the polymerizations of Glu(OBzl) NCA and Lys(Z) NCA by the poly(methyl methacrylate) initiator, the terminal amino group underwent a nucleophilic addition reaction to NCA and initiated the polymerization, yielding A-B-type block copolymers in a high yield. The same was observed in the polymerizations of Gly(OBzl) NCA and Lys(Z) NCA by the polystyrene initiator. By eliminating the protecting groups of the side chains of the polypeptide segment, the block copolymers poly(methyl methacrylate)-poly(Glu), poly(methyl methacrylate)-poly(Lys), polystyrene-poly(Glu) and polystyrene-poly(Lys) were synthesized with little side reactions. The side chain amino groups of poly(Lys) segment in the poly(methyl methacrylate)-poly(Lys) block copolymers were sulphonated or stearoylated successfully.  相似文献   

12.
Water-soluble, random copolymers containing L -methionine and N5-(3-hydroxypropyl)-L -glutamine have been prepared, fractionated, and characterized. The thermally induced helix-coil transitions of these copolymers in water have been investigated, and it has been found that incorporation of L-methionine increases the helix content of the polymers at all temperatures in the range of 0–60°C. The Zimm-Bragg parameters σ and s for the helix-coil transition in poly(L -methionine) in water were deduced from an analysis of the melting curves of the copolymers using the methods described in earlier papers.  相似文献   

13.
Plant prolyl hydroxylase recognizes poly(L-proline) II helix   总被引:5,自引:0,他引:5  
Substrate specificity of a prolyl hydroxylase from Vinca rosea suspension-cultured cells was studied using synthetic oligo(L-proline)s and their t-butyloxycarbonyl derivatives (Pron and Boc-Pron; n = 2-8) as peptidyl substrates. All peptides with a residual number of 5 or greater served as substrates in the enzyme reaction at 30 degrees C, after the preincubation of the enzyme and peptides at 0 degrees C prior to addition of cofactors and cosubstrate. Under the same conditions, the hydroxylation of Pro5 reached a plateau within 10 min, but that of Boc-Pro8 and poly(L-proline) increased linearly up to 40 min. If the preincubation temperature was raised to 30 degrees C, only Pro5 among the peptides was unable to serve as a substrate. The optimum temperature of the enzyme was 30 degrees C toward Boc-Pro8 and poly(L-proline) but it decreased to 15 degrees C using Pro5. These data suggest that the enzyme can bind Pro5 only at low temperature. Poly(L-proline) and Boc-Pron (n greater than or equal to 5) in aqueous solution are known to have a left-handed helical structure (poly(L-proline) II helix). Moreover, Pro5 was indicated as forming this helix below 10 degrees C. Accordingly, the enzyme recognizes the poly(L-proline) II helix, that is, the secondary structure of a substrate rather than the primary structure.  相似文献   

14.
The terminal amino groups of polysarcosine, poly(γ-benzyl l-glutamate), and poly(ε-benzyloxycarbonyl-l-lysine) were haloacetylated. The mixture of the terminally haloacetylated poly(α-amino acid) and styrene or methyl methacrylate was photoirradiated in the presence of Mn2(CO)10, or heated with Mo(CO)6, yielding A-B-A-type block copolymers consisting of poly(α-amino acid) (the A component) and vinyl polymer (the B component). The block copolymers were characterized, and the present investigation revealed that the thermally initiated polymerization of vinyl compounds by the trichloroacetyl poly(α-amino acid)/Mo(CO)6 system was the most suitable for the synthesis of the α-amino acid/vinyl compound block copolymers. The A-B-A type block copolymers showed higher antithrombogenicity than the corresponding homopolymers. In particular, a film of the A-B-A-type block copolymer of poly[Glu(OBzl)] and polystyrene possessed a microphase-separated structure and did not induce a conformational change of fibrinogen adsorbed, leading to a high antithrombogenicity.  相似文献   

15.
Secondary products in mycorrhizal roots of tobacco and tomato   总被引:4,自引:0,他引:4  
Colonization of the roots of various tobacco species and cultivars (Nicotiana glauca Grah., N. longiflora Cav., N. rustica L., N. tabacum L., N. tabacum L. cv. Samsun NN, N. sanderae hort. Sander ex Wats.) as well as tomato plants (Lycopersicon esculentum L. cv. Moneymaker) by the arbuscular mycorrhizal fungus Glomus intraradices Schenck and Smith resulted in the accumulation of several glycosylated C13 cyclohexenone derivatives. Eight derivatives were isolated from the mycorrhizal roots by preparative high performance liquid chromatography (HPLC) and spectroscopically identified (MS and NMR) as mono-, di- and triglucosides of 6-(9-hydroxybutyl)-1,1,5-trimethyl-4-cyclohexen-3-one and monoglucosides of 6-(9-hydroxybutyl)-1,5-dimethyl-4-cyclohexen-3-one-1-carboxylic acid and 6-(9-hydroxybutyl)-1,1-dimethyl-4-cyclohexen-3-one-5-carboxylic acid. In contrast to the induced cyclohexenone derivatives, accumulation of the coumarins scopoletin and its glucoside (scopolin) in roots of N. glauca Grah. and N. tabacum L. cv. Samsun NN, was markedly suppressed.  相似文献   

16.
In order to establish the importance of the target organ in the activation of bladder carcinogens, we compared rat liver and urothelial cell alpha-hydroxylation activities using as substrates N-nitrosobutyl(4-hydroxybutyl)amine and its metabolite N-nitrosobutyl(3-carboxypropyl)amine, two potent urinary bladder carcinogens in animals. Previous studies have shown that the production of molecular nitrogen can serve as an indicator of nitrosamine alpha-hydroxylation. The use of doubly 15N-labelled nitrosamines and the gas chromatography-mass spectrometric detection of 15N2 formed gives a measurement of the extent of this metabolic step. Various amounts of 15N-labelled substrates were incubated for 60 min at 37 degrees C with rat liver S9 preparations or urothelial cell homogenates in the presence of a NADPH generating system. Both enzyme sources metabolized 15N-labelled N-nitrosobutyl(4-hydroxybutyl)amine and N-nitrosobutyl(3-carboxypropyl)amine through the alpha-hydroxylation pathway. Using hepatic S9 fractions, 15N2 production from 15N-labelled N-nitrosobutyl(4-hydroxybutyl)amine increased from 1.69 +/- 0.02 nmol/h per mg protein (mean +/- S.E.) to 5.78 +/- 0.5 with substrate concentrations ranging between 0.55 and 5.55 mM. 15N2 produced by urothelial cell homogenates was about 40-50% that of the liver S9. 15N-labelled N-nitrosobutyl(3-carboxypropyl)amine was also metabolized through the alpha-hydroxylation pathway both by hepatic S9 and urothelial cell homogenates, though to a lesser extent. 15N2 production was about 10-times less than from 15N-labelled N-nitrosobutyl(4-hydroxybutyl)amine, but again urothelial cell 15N2 production was about 40-50% that of the liver. Treatment with phenobarbital resulted in a 2.7-fold increase in the 15N2 produced from 15N-labelled N-nitrosobutyl(4-hydroxybutyl)amine by hepatic S9. No effect was observed with urothelial cell homogenates. Acetone treatment had no effect on 15N2 production from 15N-labelled N-nitrosobutyl(4-hydroxybutyl)amine by hepatic S9, but raised 15N2 production by urothelial cell homogenates 1.8 times. Although the liver has a greater capacity than the bladder for activating the 15N-labelled nitrosamines studied, the target organ can metabolize bladder carcinogens, thus increasing the possibility of a local toxic effect. Moreover, the distribution of P-450 isozymes might be different in the bladder and this could affect the metabolism of nitrosamines reportedly formed in the human bladder in some pathological conditions.  相似文献   

17.
A series of 24 peptides Z-Gly-Xaa(R)-OH where Xaa = 15 different residues and R = H, NH2, tBu, Bzl, Trt, Mtr, and StBu were coupled with valine benzyl ester in dimethylformamide or dichloromethane at +5 degrees. The accompanying racemization was determined by analysis of the epimeric products by normal phase high-performance liquid chromatography (HPLC) for Xaa(R) = Met, Cys(StBu) and Lys(Z) and by reversed-phase HPLC after removal of benzyl-based protecting groups for Xaa(R) = Ser(tBu), Thr(tBu) and Arg(Mtr). The coupling methods examined included mixed anhydride (MxAn) at -5 degrees, and N,N'-dicyclohexylcarbodiimide (DCC), benzotriazol-1-yl-tris(dimethylamino)phosphonium hexafluorophosphate (BOP) and O-benzotriazol-1-yl-N,N,N',N'-tetramethyluroniumhexafluorophosp hate (HBTU) in the presence of 1-hydroxybenzotriazole (HOBt). Very few couplings gave stereochemically pure products. The order of sensitivity to racemization of residues depended on the method of coupling and the solvent. It varied most when comparing MxAn to HOBt-assisted reactions; it varied moderately when comparing HOBt-assisted reactions. There was less variation in comparing BOP and HBTU reactions that are initiated by the same mechanism. Leu, Nle, Phe, Asn, Lys(Z) and Asp(OBzl) are identified as the residues least sensitive to racemization. DCC-HOBt generally led to less epimerization than the other methods.  相似文献   

18.
"Stealth" nanoparticles made from polymer micelles have been widely explored as drug carriers for targeted drug delivery. High stability (i.e., low critical micelle concentration (CMC)) is required for their intravenous applications. Herein, we present a "core-surface cross-linking" concept to greatly enhance nanoparticle's stability: amphiphilic brush copolymers form core-surface cross-linked micelles (nanoparticles) (SCNs). The amphiphilic brush copolymers consisted of hydrophobic poly(epsilon-caprolactone) (PCL) and hydrophilic poly(ethylene glycol) (PEG) or poly(2-(N,N-dimethylamino)ethyl methacrylate) (PDMA) chains were synthesized by macromonomer copolymerization method and used to demonstrate this concept. The resulting SCNs were about 100 times more stable than micelles from corresponding amphiphilic block copolymers. The size and surface properties of the SCNs could be easily tailored by the copolymer's compositions.  相似文献   

19.
Poly(Glu(OBzl)-Gly)n, poly(Glu-Gly)n, poly(Gly)-(Glu(OBzl)-Gly), and poly(Gly-Glu-Gly) were synthesized from the pentachlorophenyl esters of the sequential monomer. Both of the polymers containing free glumatic-acid residues are soluble in water, as is the lower molecular weight fraction of the polytripeptides with the benzyl ester in place. Circular dichroism studies and infrared dichroism studies suggest that the 21 helix is favored for the polydipeptide with removal of the benzyl ester reducing the conformational integrity. The polytripeptide showed evidence of 31 helix in addition to the 21 form, depending on solvent. A rationale for the conformations observed is developed based on the bulkiness of the side-chain residues and conformational stabilization, in certain cases, by hydrophobic interactions between the benzyl ester groups.  相似文献   

20.
Li Y  Li Q  Li F  Zhang H  Jia L  Yu J  Fang Q  Cao A 《Biomacromolecules》2006,7(1):224-231
This study presents investigations on new approaches to novel biodegradable amphiphilic poly(L-lactide)-b-dendritic poly(L-lysine)s bearing well-defined structures. First, two new Boc-protected poly(L-lysine) dendron initiators G(2)OH 4 (generation = 2) and G(3)OH 6 (generation = 3) with hydroxyl end functional groups were efficiently derived from corresponding precursors 3 and 5 via methyl ester substitution with ethanolamine. Subsequently, two series of new diblock copolymers of poly(L-lactide)-b-dendritic Boc-protected poly(L-lysine)s (S1-S2, S3-S4) were prepared in chloroform through ring-opening copolymerization of poly(L-lactide)s with a metal-free catalyst of organic 4-(dimethylamino) pyridine (DMAP) in the presence of a corresponding new poly(L-lysine) dendron initiator. Further, molecular structures of the prepared new dendron initiators as well as those of poly(L-lactide)-b-dendritic Boc-protected poly(L-lysine)s bearing different dendron blocks and PLLA lengths were examined by means of nuclear magnetic resonance spectroscopy (NMR), gel permeation chromatography (GPC), mass spectrometry (ESI-MS, MALDI-FTMS), and thermal gravimetric analysis (TGA). The results demonstrated successful formation of the synthetic precursors, functional dendron initiators, and new diblock copolymers. In addition, the very narrow molecular weight distributions (PDI = 1.10-1.14) of these poly(L-lactide)-b-dendritic Boc-protected poly(L-lysine)s further indicated their well-defined molecular structures. After the efficient Boc-deprotection for the dendron amino groups with TFA/CH(2)Cl(2), new diblock poly(L-lactide)-b-dendritic poly(L-lysine)s bearing lipophilic PLLA and hydrophilic dendritic PLL were finally prepared. It was noteworthy that the MALDI-FTMS result showed that no appreciable intermolecular chain transesterification happened during the ROP of L-lactide catalyzed by the DMAP. Moreover, self-assembly of these new biodegradable amphiphilic copolymers in diverse solvents were also preliminarily studied.  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