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.
H Yamamoto  T Hayakawa  J T Yang 《Biopolymers》1974,13(6):1117-1125
Poly(Nδ-carbobenzoxy, Nδ-benzyl-L -ornithine) (PCBLO) was prepared by the standard NCA method. PCBLO was converted into poly(Nδ-benzyl-L -ornithine) (PBLO) through decarbobenzoxylation with hydrogen bromide. The monomer Nδ-benzyl-L -ornithine was synthesized by reacting L -ornithine with benzaldehyde, followed by hydrogenation. The conformation of the two polypeptides was studied by optical rotatory dispersion and circular dichroism. PCBLO forms a right-handed helix in helix-promoting solvents. In mixed solvents of chloroform and dichloroacetic acid (DCA) it undergoes a sharp helix–coil transition at 12% (v/v) DCA at 25°C, as compared with 36% for poly(Nδ-carbobenzoxy-L -ornithine) (PCLO). Like PCLO, the helix–coil transition is “inverse,” that is, high temperature favors the helical form. PBLO is soluble in water at pH below 7 and has a “coiled” conformation. In 88% (v/v) 1-propanol above pH (apparent) 9.6 it is completely helical. In 50% 1-propanol the transition pH (apparent) is about 7.4; this compares with a pHtr of about 10 for poly-L -ornithine in the same solvent.  相似文献   

2.
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   

3.
Poly-β-benzyl-L -aspartate (poly[Asp(OBzl)]) forms either a lefthanded α-helix, β-sheet, ω-helix, or random coil under appropriate conditions. In this paper the Raman spectra of the above poly[Asp(OBzl)] conformations are compared. The Raman active amide I line shifts from 1663 cm?1 to 1679 cm?1 upon thermal conversion of poly[Asp(OBzl)] from the α-helical to β-sheet conformation while an intense line appearing at 890 cm?1 in the spectrum of the α-helix decreases in intensity. The 890 cm?1 line also displays weak intensity when the polymer is dissolved in chloroform–dichloroacetic acid solution and therefore is converted to the random coil. This line probably arises from a skeletal vibration and is expected to be conformationally sensitive. Similar behavior in the intensity of skeletal vibrations is discussed for other polypeptides undergoing conformational transitions. The Raman spectra of two cross-β-sheet copolypeptides, poly(Ala-Gly) and poly(Ser-Gly), are examined. These sequential polypeptides are model compounds for the crystalline regions of Bombyx mori silk fibroin which forms an extensive β-sheet structure. The amide I, III, and skeletal vibrations appeared in the Raman spectra of these polypeptides at the frequencies and intensities associated with β-sheet homopolypeptides. Since the sequential copolypeptides are intermediate in complexity between the homopolypeptides and the proteins, these results indicate that Raman structure–frequency correlations obtained from homopolypeptide studies can now be applied to protein spectra with greater confidence. The perturbation scheme developed by Krimm and Abe for explaining the frequency splitting of the amide I vibrations in β-sheet polyglycine is applied to poly(L -valine), poly-(Ala-Gly), poly(Ser-Gly), and poly[Asp(OBzl)]. The value of the “unperturbed” frequency, V0, for poly[Asp(OBzl)] was significantly greater than the corresponding values for the other polypeptides. A structural origin for this difference may be displacement of adjacent hydrogen-bonded chains relative to the standard β-sheet conformation.  相似文献   

4.
The helix–coil transitions of poly(Nε-methyl, Nε-carbobenzoxy-L -lysine), poly(Nδ-methyl, Nδ-carbobenzoxy-L -ornithine), and poly(Nδ-ethyl, Nδ-carbobenzoxy-L -ornithine) in chloroform–dichloroacetic acid and their corresponding decarbobenzoxylated polypeptides in alkaline solutions were followed by optical rotation measurements. The introduction of a methyl or an ethyl group to the side chains of the carbobenzoxy derivatives of poly(L -lysine) and poly(L -ornithine) appeared to weaken the helical conformation at 25°C. The thermodynamic quantities of the three water-soluble polypeptides were calculated from the data on potentiometric titrations at several temperatures. For uncharged coil-to-helix transition, ΔH = ?370 cal/mol and ΔS = ?1.1 eu/mol for poly(Nε-methyl-L -lysine), and ΔH = ?540 cal/mol and ΔS = ?1.6 eu/mol for poly(Nδ-ethyl-L -ornithine) (all on molar residue basis). The absolute values of ΔH and ΔS dropped in the region of pH-induced transition and eventually both quantities became positive. The initiation factor σ was about 2 × 10?3, which was essentially independent of temperature. For poly(Nδ-methyl-L -ornithine) the coil-to-helix transition was not complete even when the polymer was uncharged at high pH.  相似文献   

5.
Poly(Nε-trimethyl-L -lysine), [Lys(Me3)]n, and poly(Nδ-trimethyl-L -ornithine), [Orn(Me3)]n, in sodium dodecylsulfate do not assume the β-structure or α-helix, respectively, of their parent polymers. In 0.5M Ca(ClO4)2 both [Lys(Me3)]n and [Orn(Me3)]n are aggregated and display CD spectra indicative of a regular, perhaps helical, structure. For [Lys]n and [Lys(Me3)]n, the T1 of the α-hydrogens are 0.379 and 0.230 sec, respectively, indicating greater rigidity for [Lys(Me3)]n. The CD spectrum of [Lys(Me3)]n at pH 8 is more heat resistant than that of [Lys]n. It is suggested that apolar interactions are more important in the methylated polymers than in the parent polymers.  相似文献   

6.
Poly(Nε-stearyl-L -lysine) and poly(Nε-pelargonyl-L -lysine) were synthesized both by polymerization of Nε-pelargonyl and Nε-stearyl-L -lysine NCA and by acylation of poly(L-lysine) with pelargonyl and stearyl chloride. This second route has proven to be very useful, since completely acylated polymers are obtained in almost quantitative yield, whereas the usual scheme of preparation of ε protected poly(L-lysine) cannot easily be applied due to solubility problems. Poly(Nεpelargonyl and stearyl-L -lysine) are soluble in alcohols containing linear aliphatic chains such as n-butanol and n-octanol and in mixtures of these alcohols with hydrocarbons such as n-hexane and n-heptane. Both polymers show an α-helical conformation in the above solvents, which can be disrupted upon addition of sulfuric acid. Also in the solid state, poly(Nε-stearyl-L -lysine) and poly(Nε-pelargonyl-L -lysine) show X-ray diffraction patterns typical of order structure.  相似文献   

7.
The conformation of three sequential copolypeptides, poly(L -tyrosyl-L -lysine), poly(L -tyrosyl-L -lysyl-L -lysine), and poly[L -tyrosyl-(L -lysyl)2-L -lysine] have been studied by a variety of techniques, including CD, ir spectroscopy, analytical ultracentrifugation, and x-ray diffraction. Depending upon the pH and sovent composition, poly(L -tyrosyl-L lysyl-L -lysine) and poly [L -tyrosyl-(L lysyl)2-L -lysine] can adopt either the α-helical or random-coil conformation, while poly(L -tyrosyl-L -lysine) forms either inter- or intramolecular β-structures.  相似文献   

8.
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   

9.
The hydrodynamic properties of α-helical poly(L -glutamic acid), (Glu)n in aqueous solutions and in mixtures of water with organic solvents have been interpreted in terms of the persistence length of the macromolecule. A modification of the method of Vitovskaya and Tsvetkov has been proposed in order to allow a more accurate determination of this parameter. The addition of an organic solvent increases strongly the rigidity of the helical conformation of (Glu)n. A comparison is made with some data of the literature of poly[N5-(3-hydroxy propyl)L -glutamine], [Gln(CH2)3OH]n, and poly(γ-benzyl-L -glutamate), [Glu(OBzl)]n.  相似文献   

10.
Membranes were prepared from fresh, washed human erythrocytes by hemolysis and washing with 5 mm sodium phosphate buffer (pH 7.4). The mean residue ellipticity, [θ], of erythrocyte membrane circular dichroism was altered by prostaglandin E1 or prostaglandin F at 37 °C when observed from 250 nm to 190 nm. The decrease in negativity of [θ] with 10?6m prostaglandin E1 was 12.7% at 222 nm and 17.7% at 208 nm, and with 10?6m prostaglandin F 22.5% and 34.2%, respectively (P < 0.01). Similar changes in [θ] were observed at lower concentrations of prostaglandins. No strict relationship between amount of change of [θ] and prostaglandin concentrations of 3 × 10?5m to 3 × 10?12m was evident. A persistent alteration of [θ] with prostaglandin was observed at 37 °C. Transient change of [θ] occurred at 25 °C with prostaglandin. No change of [θ] was observed at 15 or 20 °C. Buffer or palmitic acid were without effect on membrane [θ]. Phosphatidyl inositol or methyl arachidonate caused an increase in negativity of membrane spectra. The observed alterations of membrane [θ] did not arise from changes in light scattering as the OD700–OD200 of membranes was not changed by prostaglandin. Effects of prostaglandin were not dependent on light path length. The prostaglandin E1 antagonist, 7-oxa-13-prostynoic acid, at 10?7m produced no change of [θ] of membrane spectra and prevented the otherwise demonstrable effects of 10?10m prostaglandin E1 on [θ]. The decrease in negativity of [θ] at 222 nm is indicative of a decrease in ellipticity of membrane protein. These studies suggest that prostaglandins may act by inducing a conformational change in membrane protein.  相似文献   

11.
A set of large positive extrinsic CD bands ([θ]333 = 2.6 X 104 deg-cm2/decimole phosphate) in the > 300 nm region as well as diminution of the intrinsic signals (θ275) have been observed in the CD spectra of various nucleic acids complexed with the achiral compound, N-poly{α-[N-(4-pyridylethylene-4-pyridyl-N′-)α′-p-xylyl]dibromide}-4-pyridylethylene-4-pyridinium bromide, (polymer X).1,2,5 The signal changes are attributed to the binding of polymer X chromophores isogeometrically to the DNA helix in an ordered chiral arrangement. Fractionation of polymer X gives 10 well-separated oligomers. The oligomers were characterized by nmr. Their interactions with DNA have been investigated with respect to r(r = ratio of equivalents of polymer X charge/g-atoms DNA phosphorus) and n (oligomer chain length). In all cases where n ≥ 1, [θ]333 increases linearly with increasing r between 0 and 0.32, and is accompanied by a corresponding decrease in [θ]275, which becomes negative as r approaches .32. Extrinsic band intensities reveal a dependence on n up to n = 5, above which increases in nonspecific binding result in a reduction in normalized band intensities. Polymer X shows a strong preference for B-form nucleic acids and induces maximum extrinsic CD signal intensities with A-T homopolymers. Alterations in helix hydration are believed to accompany complex formation. Inversions in [θ]275 of the octamer X-poly(dA-dT) complex have been attributed to the “alternating B” conformation of poly(dA-dT).3 Similar inversions are not observed in other nucleic acid-octamer X complexes. Visible and CD spectrometry data from competition studies in the presence of the antibiotics actinomycin D (AMD), daunomycin (DM), and distamycin A (DST) are consistent with “nonclassical” intercalation as the mode of binding, and these data place the potential binding site in or near the hydrophobic region of the minor groove. Reductions in [θ]333 with increasing urea further implicate the involvement of hydrophobic interactions in the formation of an asymmetric complex. Stabilization of the helix results in all cases as evidenced by alterations in Tm; corresponding changes, however, in cooperativity are not clearly discernable. Viscosity and light-scattering data indicate no changes in molecular weight due to aggregation, and as such are not consistent with a transition to the ψ-DNA upon complex formation.  相似文献   

12.
In aqueous solutions of the amylase–iodine complex the concentration of free iodine [If]v after reaching equilibrium (or closely approximating it) is determined by the following factors: temperature, pH, concentration of iodide ions and amylose, and DP of amylose. In the present paper the role of temperature, amylose concentration, and DP has been investigated. At half-saturation of amylose by iodine, the reciprocal value of free iodine defines the equilibrium constant: 1/[If]v = K. The relation between [If]v, in normality and temperature is the following: 5 + log [If]v = ?(2.132/T) + 8.52, for DP n = 1290, 0.4 mg. amylose in 100 ml. 0.1N HCl. The value of the energy of activation Ea between 2 and 52°C. is 9.72 kcal./mole. The influence of amylose concentration [Am] on photometrically determined [If]v, at 20°C, in the range of 0.1–1.2 mg./100 ml. 0.1 N HCl for DP n = 1290 is: 5 + log [If]v = 0.209 ? 0.047 log [Am]. At [Am] = 0.6 mg. amylose/ 100 ml. 0.1 N HCl and 20°C, the value of [If]v depends on DP n as follows: 5 + log [If]v = 0.085 = + 0.222 log (104/DP n). These above equations are summarized by the relation: [If]v = exp {16.865 ? (Ea/RT)}[Am]0.047(104/DP n)0.222 ×10?5 Considering that the determination of [If]v by automatic photometric titration can be performed quickly and with appropriate reproducibility, this method is convenient for a rapid empirical and approximate determination of DP of amylose on a microscale. The iodine-binding capacity [IBC] as well as the value of λmax, have been also investigated as functions of DP n, by photometric and by amperometric titration.  相似文献   

13.
Iwao Satake  Jen Tsi Yang 《Biopolymers》1975,14(9):1841-1846
The conformational phase diagram of poly(L -lysine) (4.6 × 10?4 M, residue) in sodium dodecyl sulfate (1.6 × 10?2 M) solution was constructed from circular dichroism results at various temperatures and pH's. Poly(L -lysine)–sodium dodecyl sulfate complexes undergo a β–helix transition upon raising the pH of the solution. The transition pH tends to shift downward at elevated temperatures. No helix–β transition can be detected for poly(L -lysine) in sodium dodecyl sulfate solution (pH > 11) even after 1-hr heating at 70°C. This is in marked contrast with uncharged poly(L -lysine) solution without sodium dodecyl sulfate, which is converted into the β-form upon mild heating of the solution above 50°C.  相似文献   

14.
Abstract

Synthesis and biochemical screening against guanase of analogues of the naturally occurring guanase inhibitor azepinomycin (2) are reported. Compound e-amino-5,6,7,8,-tetrahydro-4H-imidazo[4,5-e][1,4]diazepine-5,8-dione (3) was synthesized in six steps commencing with 1-benzyl-5-nitroimidazole-4-carboxylic acid (5). Compound 3 and its synthetic precursor 3-benzyl-6-(N-benzyloxycarbonyl)amino-5,6,7,8-tetrahydro-4H-imidazo[4,5-e][1,4]diazepine-5,8-dione (12) were screened against rabbit liver guanase. Both were found to be moderate inhibitors of the enzyme with K1′s in the range of 10?4 M.  相似文献   

15.
A polypeptide having the repealing sequence (Tyr-Ala-Glu)n was synthesized by the polymerization of the N-hydroxysuccinimide ester of O-benzyl-L -tyrosyl-L -alanyl-γ-benzyl-L -glutamate, followed by the removal of the benzyl groups by means of hydrogen bromide. The main fraction obtained on gel filtration had an average molecular weight of over 60, 000, corresponding to over 500 amino acid residues per polypcptide chain. The polymer is soluble in water above pH 5.5, and precipitates on lowering the pH. The x-ray powder photographs show features of an α-helical structure. The dependence of the ultraviolet absorption spectrum, the optical rotatory dispersion, and the fluorescence of poly(Tyr-Ala-Glu) on pH, in salt-free as well as in salt-containing aqueous solutions, was compared with the corresponding properties of a copolymer containing equimolar proportions of tyrosine, alanine, and glutamic acid in a random sequence. From these measurements it was concluded that poly(Tyr-Ala-Glu ) has a helical con formation at low pH and a random coil conformation at high pH, the transition taking place at pH 6 in the absence of salt and pH II in the presence of salt. Thus, in the range pH 7 to l0. random coil-to-helix transition can be achieved by merely increasing the ionic strength. A model is proposed for the structure of the helical poly peptide which accounts for the Stability of the helical conformation by assuming hydrogen bonding between the carboxylate group of the ith glutamic acid residue and the hydroxyl group of the (i + 4 )th tyrosine residue. The complex ORD of helical poly(Tyr-Ala-Glu) is explained as being due to a superposition of the ORD of an α-helix and that of a regular array of phenolic ehroniopholes originating from the immobilization of the aromatic rings in the specific structure of the polymer.  相似文献   

16.
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   

17.
High-molecular-weight poly(0,0′-dicarbobenzoxy-L -β-3,4-dihydroxyphenyl-α-alanine) was prepared by the N-carboxyanhydride method. From the results obtained by a study of the optical rotation, nuclear magnetic resonance, and solution infrared absorption, the conformation of poly(0,0′-dicarbobenzoxy-L -β-3,4-dihydroxyphenyl-α-alanine) depended greatly on the solvent taking a right-handed helix with [θ]225 = ?13,600 ~ ?18,900 in alkyl halides, a left-handed helix with [θ]228 = 22,100 ~ 24,800 in cyclic ethers or trimethylphosphate, and a random coil structure in dichloroacetic acid, trifluoroacetic acid, or hexafluoroacetone sesquihydrate. The polypeptide underwent a right-handed helix-coil transition in chloroform/dichloroacetic acid (or trifluoroacetic acid) mixed solvents and a left-handed helix-coil transition in dioxane/dichloroacetic acid (or trifluoroacetic acid) mixed solvents. The results were compared with those of poly(0-carbobenzoxy-L -tyrosine).  相似文献   

18.
The enzymatic acylation of polar dipeptides was investigated. First, the Novozym435®-catalyzed acylation of Lys-Ser, HCl exhibiting three potential acylable sites was carried out in organic media (2-methyl-2-butanol, oleic acid) and in an ionic liquid ([Bmim]+[PF6]?). In these reactions, the chemo-selectivity of the acylation was exclusively in favour of the N?-lysine acylation and the efficiency (substrate conversion) was demonstrated to be under control of the peptide solubility. The use of [Bmim]+[PF6]? permitted to significantly improve the dipeptide solubility, and to enhance both substrates conversion and initial rates of acylation reaction. In the three reaction media used, the O-acylated derivative of the dipeptide was never detected suggesting a weak reactivity of the serine hydroxyl group due to its molecular environment and particularly to the presence of a free carboxylic group known for its electroattractor property.Last, the acylation of a natural dipeptide (carnosine), exhibiting a very low solubility in organic solvents, was also performed. Carnosine was successfully N-acylated in 2-methyl-2-butanol, and a yield of 39% was obtained when improving the substrate solubility: a better dispersibility was obtained by application of a high pressure on the reaction medium just before starting the reaction.  相似文献   

19.
K Okita  A Teramoto  H Fujita 《Biopolymers》1970,9(6):717-738
A new procedure for evaluating u and σ characterizing σ-helix-forming polypeptides in solution was derived from Nagai's theory for the helix–coil transition of such polymers. Here u is the activity for helix formation from random coil, and σ is the helix initiation parameter. The necessary data are the helical content fN at fixed solvent and temperature as a function of N, where N is the degree of polymerization of the polypeptide sample. Such data were obtained from ORD measurements on a number of fractionated samples of poly-N5-(3-hydroxypropyl)-L -glutamine (PHPG) in mixtures of water and methanol covering the complete range of composition and at various termperatures (5–40°C). When analyzed in terms of the proposed procedure, they yielded values of σ which were in the range (3.2 ± 0.6) × 10?4, substantially independent of solvent composition and temperature. These values were much larger than those obtained recently for σ of poly(β-benzyl-L -aspartate) in m-cresol and in a mixture of chloroform and DCA. The data for [η] and s0 (limiting sedimentation coefficient) as functions of molecular weight indicated that the molecular shape of PHPG in pure methanol is essentially rodlike, whereas that in pure water is not entirely randomly coiled but rather may be regarded as an interrupted helix. These indications were consistent with the results from ORD measurements. When plotted against the corresponding values of fN, the values of [η] and [s0] for PHPG in mixtures of water and methanol of various compositions and temperatures formed smooth composite curves, and we attributed these phenomena to the fact that σ of PHPG was nearly constant under these solvent conditions. Here [s0] stands for a reduced limiting sedimentation coefficient which is equal to the inverse friction factor of the solute molecule.  相似文献   

20.
The rate of conformational change of aqueous poly(α-L -lysine) solutions was measured using the electric field pulse relaxation method with conductivity detection. The relaxation time as a function of pH exhibits two maxima. One is assigned to a proton transfer reaction and the other to the helix–coil conformational transition. The helix nucleation parameter and the maximum relaxation time yield the rate constant of helix growth process (kF) according to Schwarz's kinetic theory as kF = 2 × 107 sec?1, which is comparable to that of the poly(glutamic acid) solution. The thermodynamic parameters of the helix growth process are compared with those of poly(glutamic acid).  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