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.
The random copolymerization of the N-carboxyhydrides of γ-benzyl-L -glutamate and L -valine using triethylamine as the initiator in low dielectric media reults in high-molecular-weight copolymers at low convenrson. This behavior makes it possible to apply the monomer reactivity ration theory, which was dervied for addition polymerizations, and from the use of the copolymer composition equation, the respective monomer reactivity ratios, the average and incremental copolymer compositions, and the monomer feed ratio at any conversion can be determined. A comparison of the reactivity ratios for the copolymerization of γ-benzyl-L -glutamate NCA and L -valine NCA in benzene/methylene chloride (rG = 2.1, rV = 0.6) with those obtained using dioxane (rG = 2.7, rV = 0.3) indicates that the interchain compositional heterogeneity is greater for copolymers prepared in the dioxane. For Example, at 100% conversion of the monomeric NCAs, Poly[Glu(OBzl)50Val50] prepared in dioxance has an interchain composition ranging from 74 to 0 mol % γ-benzyl-L -glutamate, whereas in benzene/methylene chloride the interchain composition of γ-benzyl-L -glutamae ranges from 65 to 0 mol %. Once the reactivity ratios are obtained for any pair of α-amino and N-carboxyanhydrides, the use of the aforementioned parameters relating to interchain composition can give insight into the compositional heterogeneity between chains as a function of conversion and provide a basis for the preparation of random α-amino acid copolymers that are homogeneous.  相似文献   

2.
R Mandel  G D Fasman 《Biopolymers》1975,14(8):1633-1649
A series of copolymers of L -lysine and L -valine [poly(L -lysinef L -valine100-f)] containing 0–13% L -valine have been studied, in 0.10M KF solution, using potentiometric titration and circular dichroism spectroscopy. Incorporation of increasing amounts of valine into the copolymers favors β-sheet formation over α-helix formation at high pH and room temperature. The titrations were analyzed using the method of Zimm and Rice and the partial free energy (ΔG0) for the coil-to-β-sheet transition for valine is estimated at 900 cal/mole at 25°C. From the temperature dependence of the free energy, the partial enthalpy, ΔH0, and entropy, ΔS0, of the transition for valine is estimated to be 854 cal/mole and 6.0 e.u., respectively. The corresponding partial thermodynamic parameters for L -lysine are in agreement with published results. The fraction of β-sheet versus pH has been calculated for poly(L -lysine86.8 L -valine13.2) at 25.0°C using the titration data; data obtained from circular dichroism spectroscopy for the same copolymer are in good accord. It is concluded from these results that L -valine is a very strong β-sheet forming amino acid. Furthermore, these results indicate that the Zimm–Rice method is applicable to transitions between the coil and β-sheet states for a polypeptide containing two different residues.  相似文献   

3.
Variation in the solvent used for the copolymerization of γ-benzyl-L -glutamate and L -valine N-carboxyanhydrides provides copolymers which have variable interchain compositions, and this variation in interchain compositional heterogeneity is reflected in the solid-state conformations of the respective copolymers. Poly[Glu(OBzl)29Val71] prepared in dioxane exhibits a β-structure, whereas a copolymer of the same average composition prepared in benzene/methylene chloride shows predominantly an β-helix conformation with a small amount of β-structure. The use of the monomer reactivity ratio permits the calculation of the average and incremental copolymer compositions at any conversion; thus, correlations between conformation and interchain compositional heterogeneity can be made. In general, copolymers prepared in dioxane show a greater distribution of chain composition and therefore permit a wider variety of conformation than copolymers prepared in benzene/methylene chloride under identical conditions.  相似文献   

4.
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   

5.
A new method, involving only gentle procedures, is described for the isolation of bovine β-casein. The optical rotatory dispersion (o.r.d.) and circular dichroism (c.d.) of the A1 variant, so isolated, are determined in the temperature range 2–60°C, at pH 6.9 (25°C) and I = 0.012 mol l?1. Conformational analyses are made of the o.r.d. and c.d. results using the reference c.d. spectra of Brahms and Brahms, the Kronig—Kramers transform, and also of the c.d. results by the method of Provencher and Glöckner. The conformations obtained are compared with those predicted for the amino acid sequence by the methods of Chou and Fasman and of Lim. It is concluded that β-casein contains 9 ± 2% α-helical structure and 25 ± 6% β-sheet at 2°C. A structural interpretation is proposed for the effect of increase of temperature on the o.r.d. and c.d., involving an increase in the proportion of β-sheet at the expense of aperiodic structure.  相似文献   

6.
We have analyzed the hemoglobins of a young German patient with β-thalassemia intermedia and of his immediate family and included in these studies an evaluation of possible nucleotide changes in the β-globin through sequencing of amplified DNA. One chromosome of the propositus and one of his father's carried the GTGGGG mutation at codon 126 leading to the synthesis of Hb Dhoburi or α2β2126(H4)Val→Gly; this variant is slightly unstable and is associated with mild thalassemic features. His second chromosome and one of his mother's had the common IVS-I-5 (G→C) mutation that leads to a rather severe β+-thalassemia and the GTGATG mutation at codon 18, resulting in the replacement of a valine residue by a methionine residue. This newly discovered β-chain variant, named Hb Baden, was present for only 2–3% in both the patient and his mother. This low amount results from a decreased splicing of RNA at the donor splice-site of the first intron that is nearly completely deactivated by the IVS-I-5 (G→C) thalassemic mutation. The chromosome with the codon 18 (GTGATG) and the IVS-I-5 (G→C) mutations has thus far been found only in this German family; analysis of 51 chromosomes from patients with the IVS-I-5 (G→C) mutation living in different countries failed to detect the codon 18 (GTGATG) change.  相似文献   

7.
Copolymers of L -lysine and L -isoleucine [poly(L -Lysf,L -Val1 ? f)] containing 4–15% isoleucine were investigated using potentiometric titration and circular dichroism (CD) spectroscopy. With increasing isoleucine content, β-sheet formation is favored over α-helix formation at high pH and room temperature. The fraction of β-sheet present, as a function of pH, calculated from titrations of poly(L -Lys85.2,L -Ile14.8), agreed well with data obtained from CD studies for the same copolymer. Thermodynamic parameters were determined from titrations using the method of Zimm and Rice; the partial free energy (ΔG°C → β) at 25° for the coil-to-β-sheet transition for isoleucine was estimated to be ?515 cal/mol; from the temperature dependence of free energy, the partial entropy (ΔS°cβ), and the partial free enthalpy (ΔH°c → β) of the coil → β transition for isoleucine is estimated to be 2.6 e.u. and 260 cal/mol, respectively. The partial thermodynamic parameters obtained for lysine are in good agreement with literature values. It is concluded from these studies that isoleucine has a very high potential for a β-sheet formation.  相似文献   

8.
The packing of α-helices and β-sheets in six αβ proteins (e.g. flavodoxin) has been analysed. The results provide the basis for a computer algorithm to predict the tertiary structure of an αβ protein from its amino acid sequence and actual assignment of secondary structure.The packing of an individual α-helix against a β-sheet generally involves two adjacent ± 4 rows of non-polar residues on the α-helix at the positions i, i + 4, i + 8, i + 1, i + 5, i + 9. The pattern of interacting β-sheet residues results from the twisted nature of the sheet surface and the attendant rotation of the side-chains. At a more detailed level, four of the α-helical residues (i + 1, i + 4, i + 5 and i + 8) form a diamond that surrounds one particular β-sheet residue, generally isoleucine, leucine or valine. In general, the α-helix sits 10 Å above the sheet and lies parallel to the strand direction.The prediction follows a combinational approach. First, a list of possible β-sheet structures (106 to 1014) is constructed by the generation of all β-sheet topologies and β-strand alignments. This list is reduced by constraints on topology and the location of non-polar residues to mediate the sheet/helix packing, and then rank-ordered on the extent of hydrogen bonding. This algorithm was uniformly applied to 16 αβ domains in 13 proteins. For every structure, one member of the reduced list was close to the crystal structure; the root-mean-square deviation between equivalenced Cα atoms averaged 5.6 Å for 100 residues. For the αβ proteins with pure parallel β-sheets, the total number of structures comparable to or better than the native in terms of hydrogen bonds was between 1 and 148. For proteins with mixed β-sheets, the worst case is glyceraldehyde-3-phosphate dehydrogenase, where as many as 3800 structures would have to be sampled. The evolutionary significance of these results as well as the potential use of a combinatorial approach to the protein folding problem are discussed.  相似文献   

9.
The accessible inclusion sites of insoluble copolymers containing β-cyclodextrin (β-CD) were studied in aqueous solutions by measuring the absorbance changes (decolourization) of phenolphthalein (phth) at pH 10.5. The various copolymers were reacted at different β-CD:crosslinker mole ratios with five individual types of crosslinker agents (epichlorohydrin (EP), sebacoyl chloride (SCL), terephthaloyl chloride (TCL), glutaraldehyde (GLU), and poly(acrylic) acid (PAA), respectively). The decolourization provided estimates of the 1:1 binding constants (K1) for the β-CD monomer/phth complex. Comparable values of K1 were measured for copolymer/phth complexes with highly accessible β-CD inclusion sites as compared with the 1:1 β-CD/phth complex. The surface accessibility of the β-CD inclusion binding sites for the polymers ranged from ∼10 to 72%. The observed variability of the inclusion sites was attributed to: (i) steric effects in the annular hydroxyl region of β-CD, (ii) the degree of crosslinking of the copolymer and (iii) the accessibility of the micropore sites within the copolymers. The Gibbs free energy (ΔG°) and site occupancy (θ) of phth adsorbed to the copolymer materials was estimated independently using the Sips isotherm model. The ΔG° values ranged between −27.6 and −30.9 kJ mol−1 for the copolymers and are in close agreement with the value for the 1:1 β-CD/phth complexes (ΔG° = −27 kJ mol−1) in aqueous solution.  相似文献   

10.
13C n.m.r. CP/MAS spectra (50.3 and 75.4 MHz) of solid poly(l-lleucines) and poly(d-norvalines) measured with suitable acquisition parameters allow quantification of the composition of the secondary structure. The optimum acquisition parameters were found by systematic variation of the contact time by means of samples containing 5?0% α-helix structure. The polypeptides were prepared by primary or tertiary amine-initiated polymerizations of the corresponding amino acid NCAs and the average degrees of polymerization (DP) were determined by 1H n.m.r. endgroup analysis. The mole fraction of α-helices increases with increasing DP; it depends on the nature of the solvent and to a lesser degree on the polymerization temperature. When prepared under identical conditions, poly(d-norvaline) samples contain more β-sheet structure than poly(l-leucine. Reprecipitation increases the α-helix content, demonstrating that a part of the original β-sheet structure is thermodynamically unstable. The presence of oligomers of DP ?10 is mainly responsible for the thermodynamically stable part of the β-sheet structure. The chain growth mechanism is discussed.  相似文献   

11.
Chicken erythrocyte chromatin was partially digested with micrococcal nuclease and separated into multimeric subunit fractions by gel permeation chromatography. The fractions were characterized by their Svedberg constant, diffusion coefficient, circular dichroism, and electrophoresis pattern of the extracted DNA. The molecular weight dependence of the sedimentation coefficient was found to be S20,w = .011 × M.554. The molecular weight dependence of rmffo is best represented in the Kirkwood theory by either a helical superstructure or a flexible coil withattractiveinteractions between nucleosome units. The dimer calculations of ffo suggest that the core particles are separated by spacer regions which contribute up to ~20% of the frictional properties of the molecule.  相似文献   

12.
Nine variant specific surface antigens were purified from clones of Trypanosoma equiperdum and characterized by amino acid analysis, isoelectric focusing and circular dichroism. The molecules showed extensive differences in their isoelectric points, and by comparison with the corresponding amino acid compositions, this variation seemed to be due to different amide contents. Circular dichroism data allowed one to divide the molecules into 4 groups according to their respective percentages inα-helical and β-sheet structure.  相似文献   

13.
Hydrophilic water-insoluble gels suitable for affinity chromatography of lectins have been prepared by copolymerization of acrylamide, N,N′-methylene bisacrylamide and alkenyl 1-thioglycosides. Water-soluble copolymers of analogous type have been obtained by omitting the cross-linking agent, N,N′-methylene bisacrylamide.In affinity chromatography of the Ricinus communis lectin it could be shown that the capacity for the lectin of the water insoluble copolymers was more than four times higher in copolymers having the S-β-D-galactosyl ligand attached through a methylene bridge than in derivatives with a nonamethylene spacer.None of the insoluble S-β-D-glycosyl copolymers prepared could be shown usable as affinity adsorbent for glycosidases though the corresponding soluble copolymers inhibited the activity of the enzymes.  相似文献   

14.
By copolymerization of acrylamide and allyl glycosides of various sugars, O-glycosyl derivatives of polyacrylamide copolymers were prepared. The sugar content of the copolymers can be varied in the range 0–40%, their sedimentation coefficient shows the values of 2.5–2.7 S; the molecular weight of an O-α-d-mannopyranosyl polyacrylamide copolymer (29% mannose, s20,w0 = 2.9 S) was estimated as 44 500. Copolymers with incorporated glycosyl residues interacting specifically with lectins yield precipitates with them upon immunodiffusion in cellulose acetate. The quantitative precipitin curves obtained with these copolymers are similar to those produced by quantitative precipitation of lectins with natural polysaccharides. The copolymers may serve as model substances of natural polysaccharides.  相似文献   

15.
In normal human subjects under basal conditions, we have reported that molar concentrations of immunoreactive β-lipotropin (IR-β-LPH) are approximately threefold greater than those of IR-β-endorphin (β-Ep). Following acute stimulation, there is a further two- to threefold disproportionate rise in plasma concentrations of IR-β-LPH as compared to those of IR-β-Ep. To begin to assess the possible factors involved in such altered IR-β-LPH/IR-β-Ep ratios in plasma, the metabolic clearance rate (MCR), volume of distribution (Vd), fractional rate of disappearance (Kd), and half-life (t12) of these peptides were determined by means of bolus injection of highly purified human β-LPH and synthetic human β-Ep in normal human subjects. β-Ep was found to have an MCR and a Kd greater than that of β-LPH, and a shorter t12. These differences, however, although they may in part be contributory, cannot solely account for the greater ratio of IR-β-LPH to IR-β-Ep in plasma, or for the disproportionate rise in plasma concentrations of these peptides after acute stimulation.  相似文献   

16.
A β-sheet conformation is predicted at the N-terminal of β chains in sickle cell hemoglobin (Hb S) as a result of the β6 Glu → Val mutation. Since Glu is the weakest and Val is the strongest β-sheet former in the predictive method of Chou and Fasman [Biochemistry 13, 211, 222 (1974)], such a substitution greatly increases the β-sheet potential in the β 1–6 region. The similarity in the concentration and temperature dependence of Hb S gelation to β-sheet formation in polyamino acids suggest that a common aggregation mechanism may be involved. Conditions to cause a β → α trans-formation at the β 1–6 region of Hb S is discussed relative to the treatment of sickle cell disease.  相似文献   

17.
Two prototype triblock (ABA) copolymers of poly[(γ-benzyl-l-glutamate) x (butadiene/acrylonitrile)y (γ-benzyl-l-glutamate)x] have been synthesized and characterized. They were prepared by reacting a primary amine capped butadiene/acrylonitrile (ATBN) polymer with the N-carboxy anhydride of γ-benzyl-l-glutamate. The copolymers were ~38 000 (copolymer 1) and 74 000 (copolymer II) molecular weight. X-ray diffraction and Fourier Transform infrared spectroscopy of films cast from dioxane (preferential for PBLG) and chloroform (non-preferential) show the benzyl glutamate segments to be predominantly α-helical and disordered α-helical, respectively. Electron microscopy of osmium tetroxide strained films cast from dioxane revealed lamellar domain formation indicative of phase separation. The midblock butadiene layers were ~150 Å thick while the alternating benzyl glutamate layers were 300 and 500 Å thick for copolymers I and II, respectively. Films cast from chloroform exhibit a nearly homogeneous morphology, indicative of considerable phase mixing. Dynamic mechanical spectroscopy of the copolymers also revealed a dependence on morphology. The side chain transition of the benzyl glutamate appeared as a single peak when the copolymers were cast from dioxane and a double peak when the copolymers were cast from chloroform.  相似文献   

18.
(1→3)-β-D-Glucans of various degrees of polymerization were prepared by degradation of a gel-forming D-glucan with formic acid. The degraded D-glucans were separated into a water-soluble fraction (soluble D-glucan) and an insoluble fraction (insoluble D-glucan). Both D-glucans were further fractionated. The optical rotation including determination of the o.r.d. curves of the fractions and of the original gel-forming D-glucans was measured at various sodium hydroxide concentrations (0–5M). The results indicate that (1→3)-β-D-glucans of DPn below ca. 25 (the soluble D-glucan) took a disordered form in both neutral and alkaline solutions, whereas the D-glucans of higher DPn (the insoluble and the original D-glucans) took an ordered structure in dilute alkaline solution (0.1M). The proportion of ordered structure in the insoluble D-glucan increases with DPn to attain a maximum value at a DPn of around 200; this may be the lower limit of DPn to permit gel formation in neutral media. The formation of complexes with Congo Red in alkaline solutions by the soluble and the insoluble D-glucans supports the same conclusions.  相似文献   

19.
A highly purified preparation of mouse β2-microglobulin has been obtained from sodium thiocyanate extracts of liver cell membranes of AJ strain mice and the amino acid sequence has been partially determined. The first 40 residues have been assigned except for position 34. In the sequence, the mouse protein differs only at 9 positions from human β2-microglobulin at 8 positions from the dog homologue and at 11 positions from the rabbit homologue. The sequence has also a homology to the constant regions of mouse γG2a, most closely to the CH3 region. These data support the conclusion that the mouse protein is indeed the mouse homologue of human β2-microglobulin.  相似文献   

20.
Following incubation with [3H]-PGF, 73–91% of the 3H activity accumulated by rabbit uterus, choroid plexus or anterior uvea was shown to remain associated with PGF on two different chromatographic systems. The tissue to medium ratios, calculated on the basis of chromatographically identified [3H]-PGF, were greater than unity (2.3–10.4) for all three tissues and the extracted 3H activity could be effectively accumulated by these tissues for a second time. Under conditions when 85% of authentic [3H]-PGF and only 8% of [3H]-15-keto-PGF was adsorbed on rabbit anti-PGF serum, 60–75% of the extracted 3H was adsorbed onto the antiserum. Following incubation with a mixture of 5,6-[3H]-PGE1 and 2-[14C]-PGE1, the anterior uvea and the uterus showed similar TM ratios for 3H and 14C and the 3H14C ratios were essentially constant in their respective homogenates, extracts and chromatographic fractions, indicating insignificant β-oxidation of the accumulated PGE1. In the case of the kidney cortex, a substantial fraction of the accumulated 14C did not extract as a PG presumably as a result of β-oxidation. It is concluded that metabolic alteration of the accumulated PG molecule does occur in some tissues, but such chemical alterations are not an integral part of the PG accumulative process. These results are consistent with the concept that some vertebrate tissues can accumulate PGs against a concentration gradient by an active transport mechanism.  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