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.
Electric dichroism studies on poly(γ-benxyl L -glutamate) show that the addition of small amounts (up to 5%) of trifluoroacetic acid causes complete disappearance of dichroism in contrast to the electric birefringence, which drops to an observable plateau at 20% of its initial value. This loss of dichroism cannot be explained simply by a decrease in the effective dipole moment of the benzyl ester side chains, and the nature of this interaction is explored by nuclear magnetic resonance and infrared spectral studies. Theories of the helix-coil transition which require an initial protonation of the helix backbone to form a more flexible rod consisting of helical segments interrupted by regions of random coil are shown to be inadequate to describe the changes in poly(γ-benzyl L -glutamate) effected by small amounts of trifluoroacetic acid. Rather, trifluoroacetic acid in small amounts interacts with the ester carbonyl oxygen in the side chain, either by hydrogen bonding or protonation, before there is evidence of any effect on the backbone or of loss of helix content.  相似文献   

2.
A Roig  M Cortijo 《Biopolymers》1971,10(2):321-328
The helix–coil transition in poly(γ-benzyl L -glutamate-co-γ-methyl L -glutamate) copolypeptides was studied experimentally in nonaqueous solvents and the results compared with theory. It is found that the transition can be described by the same theory as for the homopolypeptides, but the corresponding parameters are not related in the expected way to those of the homopolymers, due to the effect of the side chains on the stability of the helix, which is not taken into account explicitly by the theory.  相似文献   

3.
Even though poly(L -valine) and poly(L -isoleucine) both contain residues that are branched at their β-carbon atoms, they exhibit a different behavior of their Zimm-Bragg helix-growth parameter s in aqueous solution. This quantity increases with temperature for poly(L -valine) but decreases for poly(L -isoleucine). The origin of this behavioral difference was examined by computing theoretical values of s versus temperature from interatomic interaction energies, taking solvent (hydrophobic and hydrophilic) effects into account. The calculated s versus temperature curves for both homopolymers are consistent with the observed experimental behavior. The two homopolymers behave differently because of differences in the change in the number of hydration-shell water molecules accompanying their helix–coil transitions. The larger isoleucine side chains are more crowded together in both the α-helical and coil forms than are those of valine. Therefore, there is a smaller change in hydration of the isoleucine side chains compared to that of the valine side chains in the helix–coil transition. By analyzing the effects of hydration on the s versus temperature curves, it is possible to account also for the experimental curve for poly(L -leucine), which exhibits an intermediate behavior between those for poly(L -valine) and poly(L -isoleucine).  相似文献   

4.
The thermal helix–coil transition of poly(γ-benzyl L -glutamate-co-ε-carbobenzoxy-L -lysine) copolypeptides was studied in solvent mixtures of different compositions. The cooperativity parameter v changes linearly with polymer (and solvent) composition, whereas the heat of the transition shows a very pronounced minimum as a function of polymer composition. This minimum cannot be due only or mainly to the solvent changes and must be attributed to the effect on the transition of the side chains of the polypeptides.  相似文献   

5.
Light scattering of multichain poly-α-amino acids was studied in dimethylformamide (DMF). The polymers consisted of a backbone of poly-L -lysine of degree of polymerization n with side chains of benzyl L -glutamate and benzyl L -aspartate of degree of polymerization, m, on each ε-amino group. The backbone length n is known and m is obtained by amino acid analysis. The results on a series of such materials confirm this structure and show that the molecules are dissolved in highly compacted conformations. It was found that DMF is a poor solvent for these polymers. In the case of the higher molecular weight polymers, the solutions initially were not molecularly disperse. The aggregates were resistant to dilution in the experimental range. Mild heat treatment, however, disaggregated the solutions irreversibly, and the light-scattering data indicated that a structural rearrangement of the molecules had occurred.  相似文献   

6.
It has already been show that the helix senses of poly(β-benzyl L -aspartate) and poly(β-methyl L -aspartate) are left-handed, while the poly esters of n-propyl, isopropyl, n-butyl, and phenethyl L -asparate are all right-handed. The effect of changes in helix sense from the left-handed to the right-handed α-helical form on the infrared spectra of copolymers of benzyl L -aspartate with ethyl, n-butyl, isopropyl, n-propyl, and phenethyl L -aspartate have been studied. Those show that for the right-handed helical form the amide band frequencies fall within the range given by Elliott,7 while for the left-handed form the frequencies are higher. The frequency ranges for the two helix senses are given and have been used to show that poly (β-n-propyl L -aspartate) in chloroform solution undergoes a transition from the right-handed to the left-handed helix form on heating. Polarized infrared studies of the different copolymers show that the disposition of the side chain ester groups is different for the two forms. Although methyl L -aspartate forms a left-handed α-helix similar to benzyl L -aspartate, the introduction of methyl L -aspartate residues into poly (β-benzyl L -aspartate) prevents the formation of the ω-helix. The factors involved in the formation of this helix form are discussed.  相似文献   

7.
The conformational and binding properties towards Cu(II) and Ni(II) ions of Gly-Gly-His derivatives of poly(l-lysine) have been investigated mainly using circular dichroism (c.d.) spectroscopy. These derivatized polymers can be considered macromolecular analogues of the Cu(II) and Ni(II) binding site of human serum albumin. It has been shown that modification up to 53% of the ε-amino groups of lysine side chains by covalent binding of the tripeptide unit Gly-Gly-His does not induce appreciable alteration of the α-helix forming tendency of the polylysine backbone. The derivatized polymers exhibit strong affinity towards Cu(II) and Ni(II) ions. At neutral pH, complexes are formed in which each tripeptide chelating unit is linked to one metal ion. The spectral characteristics in the visible absorption region are consistent with a square planar geometry of the complexes, with deprotonated peptide groups and one imidazole nitrogen in the coordination sphere of the ion. C.d. measurements in the far u.v. indicate that complex formation in the side chains causes an increase of ordered structure of the peptide backbone at neutral pH. This fact is interpreted in terms of a reduced electrostatic repulsion among side chains due to charge neutralization in the tripeptide units linked to metal ions.  相似文献   

8.
The conformational and binding properties towards Cu(II) and Ni(II) ions of Gly-Gly-His derivatives of poly(l-lysine) have been investigated mainly using circular dichroism (c.d.) spectroscopy. These derivatized polymers can be considered macromolecular analogues of the Cu(II) and Ni(II) binding site of human serum albumin. It has been shown that modification up to 53% of the ε-amino groups of lysine side chains by covalent binding of the tripeptide unit Gly-Gly-His does not induce appreciable alteration of the α-helix forming tendency of the polylysine backbone. The derivatized polymers exhibit strong affinity towards Cu(II) and Ni(II) ions. At neutral pH, complexes are formed in which each tripeptide chelating unit is linked to one metal ion. The spectral characteristics in the visible absorption region are consistent with a square planar geometry of the complexes, with deprotonated peptide groups and one imidazole nitrogen in the coordination sphere of the ion. C.d. measurements in the far u.v. indicate that complex formation in the side chains causes an increase of ordered structure of the peptide backbone at neutral pH. This fact is interpreted in terms of a reduced electrostatic repulsion among side chains due to charge neutralization in the tripeptide units linked to metal ions.  相似文献   

9.
The interaction between bovine serum albumin (BSA) and the anionic graft copolymers poly(sodium acrylate-co-sodium 2-acrylamido-2-methyl-1-propanesulfonate)-graft-poly(N,N-dimethylacrylamide) (P(NaA-co-NaAMPS)-g-PDMAMx) was investigated within the acid pH region, 2 < or = pH < or = 7. The weight percentage, x, of the poly(N,N-dimethylacrylamide) (PDMAM) side chains varied from 0 up to 75% (w:w). When BSA and P(NaA-co-NaAMPS)-g-PDMAMx are oppositely charged, i.e., when pH is lower than the isoelectric point of BSA, the two macromolecules associate through Coulombic attractions. When the anionic graft copolymer is rich enough to the nonionic PDMAM side chains, x > or = 50% w:w, the associative phase separation is practically prevented, as revealed by the turbidimetric study of the BSA/P(NaA-co-NaAMPS)-g-PDMAMx mixtures in aqueous solution vs pH. In addition, the viscosity measurements support the formation through a charge neutralization process of a rather compact protein-polyelectrolyte complex stabilized by the hydrophilic PDMAM side chains grafted onto the anionic copolymer backbone.  相似文献   

10.
Monte Carlo (MC) simulations have been used to study the structure of an intermediate thermal phase of poly(alpha-octadecyl gamma,D-glutamate). This is a comblike poly(gamma-peptide) able to adopt a biphasic structure that has been described as a layered arrangement of backbone helical rods immersed in a paraffinic pool of polymethylene side chains. Simulations were performed at two different temperatures (348 and 363 K), both of them above the melting point of the paraffinic phase, using the configurational bias MC algorithm. Results indicate that layers are constituted by a side-by-side packing of 17/5 helices. The organization of the interlayer paraffinic region is described in atomistic terms by examining the torsional angles and the end-to-end distances for the octadecyl side chains. Comparison with previously reported comblike poly(beta-peptide)s revealed significant differences in the organization of the alkyl side chains.  相似文献   

11.
The α-helical from of poly(L -glutamic acid) [α-poly(Glu)] gives rise to the same amide I and III lines as α-poly(γ-benzyl-L -glutamate) at 1652 and 1296 cm?1, respectively. The latter is a superposition of the amide III line near 1290 cm?1 and a line deu to vibrational made of CH2 groups of the side chain near 1300 cm?1. A line at 924 cm?1 is tentatively identified as characteristics of α-poly(Glu). Both the β1- and β2- forms of poly(Glu) give rise to characteristic of β-amide. III frequencies that are similar because of their similar backbone structures. Differences in the conformations of their side chains and in the environments of the backbone are reflected in the region 800–1200 cm?1 and in the amide I. A line at 1042 cm?1 and a pair at 1021 and 1059 cm?1 are tentatively assigned as characteristic of β1-poly(Glu) and β2-poly(Glu), respectively. The α-β2 transition in poly(L -Glu78L -Val22) is shown by the appearance of all the β2-characteristic lines in the thermally transformed sample. The same features observed in poly(L -Glu95L -Val5) also indicate that the α-β2 transition of poly(Glu) is facilitated by the presence of L -valine and that the content of L -valine is not critical for this purpose. Investigation of the Raman spectra of the calcium, strontium, barium and sodium slats of poly(Glu) shows that these salts, under the conditions of preparation used, all the have random-coil conformations.  相似文献   

12.
Peters D  Peters J 《Biopolymers》2001,59(6):402-410
The pseudomolecule approach to the structure of globular proteins in which a small number of water molecules are incorporated into the "molecule" is tested again by comparing the ribbon of hydrogen bonds in two proteins, ribonuclease F1 and T1. These two molecules are 59% homologous and have the same backbone conformation both globally and locally. The two ribbons of hydrogen bonds that cover the whole of the backbone are conserved with an accuracy of some 95% providing that allowance is made for the intrusion into one of the pair of such extra factors as the presence of adducts or metal ions, the insertions and the absence of a few water molecules from one of the x-ray data sets. Without these corrections, the conservation of the ribbon is some 85%. There are 35 conserved hydrogen-bonding residues, nearly all of which show many unions to the backbone or interactions with the active site. There are 36 point mutations that involve one or two hydrogen-bonding side chains and nearly all of these have either none or one hydrogen bond to the backbone. These are minor contributors to the ribbon of hydrogen bonds. Of the 71 residues involved in these two categories, all but six fit into the pseudomolecular picture of the structure of globular proteins. The remaining 30 residues almost all contain conserved hydrocarbon side chains that may have a second order effect on the structure through their space filling effects.  相似文献   

13.
The L intermediate in the proton-motive photocycle of bacteriorhodopsin is the starting state for the first proton transfer, from the Schiff base to Asp85, in the formation of the M intermediate. Previous FTIR studies of L have identified unique vibration bands caused by the perturbation of several polar amino acid side chains and several internal water molecules located on the cytoplasmic side of the retinylidene chromophore. In the present FTIR study we describe spectral features of the L intermediate in D(2)O in the frequency region which includes the N-D stretching vibrations of the backbone amides. We show that a broad band in the 2220-2080 cm(-1) region appears in L. By use of appropriate (15)N labeling and mutants, the lower frequency side of this band in L is assigned to the amides of Lys216 and Gly220. These amides are coupled to each other, and interact with Thr46 and Val49 in helix B and Asp96 in helix C via weakly H-bonding water molecules that exhibit O-D stretching vibrations at 2621 and 2605 cm(-1). These water molecules are part of a hydrogen-bonded network characteristic of L which includes other water molecules located closer to the chromophore that exhibit an O-D stretching vibration at 2589 cm(-1). This structure, extending from the Schiff base to the internal proton donor Asp96, stabilizes L and affects the L-to-M transition.  相似文献   

14.
Pectin with [alpha]D(20) +192 degrees (c 0.1; water), named comaruman, was isolated from marsh cinquefoil Comarum palustre L., which is widespread in the European North. The sugar chain of comaruman contains residues of D-galacturonic acid (64%), D-galactose (13%), L-rhamnose (12%), L-arabinose (6%), and trace amounts of xylose and glucose. Partial acid hydrolysis and digestion with pectinase demonstrated that comaruman composed of the backbone comprised regions of linear alpha-1,4-D-galactopyranosyl uronan interconnected by numerous residues of alpha-1,2-L-rhamnopyranose. In addition to the backbone (core of the macromolecule), ramified regions are involved in comaruman and comprise alpha-2,4-L-rhamno-alpha-4-D-galacturonan with side chains consisting mainly of beta-1,4-linked residues of D-galactopyranose. The ramified region contains additionally residues of 5-O-substituted arabinofuranose and 3- and 6-O-substituted galactopyranose. The present 3,4- and 4,6-di-O-substituted residues of galactopyranose appear to be branching points of the side chains. Some galactopyranose residues were found to occupy the terminal positions of the side chains or appeared to be single sugar residues attached to the side chains. Methylation analysis data indicated that comaruman contains residues of terminal, 3- and 3,4-di-O-substituted galactopyranosyl uronic acid, which appeared to be constituents of the side chains, and the latter represented additionally branching points of the backbone.  相似文献   

15.
A new series of linear and permanently charged poly(amidoammonium) salts were synthesized in order to investigate the influence of their ionic and hydrophobic contents on both the cytotoxicity and the transfection mediated by polycation-DNA complexes. The poly(amidoammonium) salts were prepared by chemical modification of a parent poly(amidoamine) containing two tertiary amino groups per structural unit: one incorporated into the main chain and the other fixed at the end of a short bismethylene spacer. The permanent charges were introduced through a quaternization reaction involving iodomethane or 1-iodododecane as an alkylating agent. Under appropriate conditions, the methylation reaction was found to be regioselective, allowing the quaternization of either the side chains or both the side chains and the backbone. Under physiological salt conditions (150 mM NaCl), all of the poly(amidoammonium) salts self-assembled with DNA to form complexes. High proportions of highly quaternized polycation provided better defined morphology to the polycation-DNA complexes. Complexes formed from unquaternized polycation were less cytotoxic than branched poly(ethyleneimine) (25 kDa). At high polycation-DNA weight ratios, the introduction of permanent charges generated a significant increase in the cytotoxicity, but no patent correlation could be established with the amount and the position of the permanent charges. Only complexes formed from polycations with quaternized backbone were able to generate significant gene expression, which was putatively attributed to a better defined toroidal-like morphology together with a higher stability, as suggested by zeta potential measurements. The incorporation of dodecane side chains on highly charged polycations severely amplified the cytotoxicity so that, in return, the transfection level was dramatically affected.  相似文献   

16.
Photochromic polypeptides, with 16 to 56% azobenzene groups in the side chains, have been prepared by reaction of poly(L -glutamic acid) with p-aminozaobenzene, both in the presence of dicyclohexyl carbodiimide/N-hydroxybenzotriazole and of pivaloyl chloride. Analogous modification reactions carried out on poly(L -aspartic acid) were unsuccessful owing to the formation of N-succinimide rings. In trimethylphosphate, all the azopolypeptides exhibit the α-helix CD pattern. Irradiation produces the trans-to-cis isomerization of the azo side chains, but does not induce any variations of the backbone conformation. In water, the CD spectra indicate the presence of appreciable amounts of α helix in 16 and 21% mol azo-containing poly-(L glutamates), while a β structure is present in a 36% mol azopolypeptide. Light produces conformational changes of the polypeptide conformation which are completely reversed in the dark. The extent and kind of photobehavior depend on the azo content and the pH value at which irradiation is carried out. The light-induced effects are discussed on the basis of the pH-induced order-disorder conformational transitions. In fact, the pK values and the transition curves of the dark-adapted samples were found to be different from those of the irradiated ones.  相似文献   

17.
B R Gelin  M Karplus 《Biochemistry》1979,18(7):1256-1268
Side-chain torsional potentials in the bovine pancreatic trypsin inhibitor are calculated from empirical energy functions by use of the known X-ray structure of the protein and the rigid-geometry mapping technique. The potentials are analyzed to determine the roles and relative importance of contributions from the dipeptide backbone, the protein, and the crystalline environment of solvent and other protein molecules. The structural characteristics of the side chains determine two major patterns of energy surfaces, E(X1,X2): a gamma-branched pattern and a pattern for longer, straight side chains (Arg, Lys, Glu, and Met). Most of the dipeptide potential curves and surfaces have a local minimum corresponding to the side-chain torsional angles in the X-ray structure. Addition of the protein forces sharpens and/or selects from these minima, providing very good agreement with the experimental conformation for most side chains at the surface or in the core of the protein. Inclusion of the crystalline environment produces still better results, especially for the side chains extending away from the protein. The results are discussed in terms of the details of the interactions due to the surrounding, calculated solvent-accessibility figures and the temperature factors derived from the crystallographic refinement of the pancreatic trypsin inhibitor.  相似文献   

18.
A study of the influence of the spiropyran to merocyanine ring opening on a model of poly(spiropyran-L -glutamate) as implied by the experimental data (T. M. Cooper, K. A. Obermeier, L. V. Natarajan, and R. L. Crane (1992) Photochemistry and Photobiology, 55, 1–7) is presented. The individual chromophore is studied by the AM1 semiempirical approach, while molecular mechanics and dynamics calculations are employed in the analysis of the poly(spiropyran-L -glutamate) model. It is shown that the α-helical secondary structure is less conserved in the polypeptide substituted with the merocyanine form of the chromophore. In particular, larger side-chain flexibility, increased backbone hydrogen-bond lengths, as well as a larger helix bending are calculated. Furthermore, a random conformational minimization calculation finds the intrinsic behavior of the spiropyran molecular system as being more of a helix “maker” than its merocyanine analogue. The interactions of the chromophore substituent with other side chains prove, in part, that an early event in the decay of the α-helical structure is the formation of hydrogen bonds between the carboxylic acid groups and the merocyanine oxygens. The results lend support to the experimental observation that the merocyanine group destabilizes the α-helical framework of the polypeptide, thus possibly allowing the entry of solvent molecules into the α-helical core, while spiropyran in its closed form shields it from the solvent. © 1992 John Wiley & Sons, Inc.  相似文献   

19.
An alcoholysis method is described for the modification of high molecular weight poly(β-benzyl L -asparatate); by this method the benzyl groups in the polypeptide have been replaced by methyl, ethyl, isopropyl, n-propyl, and phenethyl groups to give a series of copolymers of each of the corresponding aspartate esters with benzyl L -aspartate. By repeating the reactions, replacement of better than 99% has been achieved in some cases to give in effect the homopolymer. Optical rotatory dispersion studies show that of all the systems studied only poly(β-methyl L -aspartate) has the left-handed helix sense, the others are right-handed. It is shown further that the helix sense is not an intrinsic property of the nature of the aspartate side chain. Raising the temperature of chloroform solutions of the right-handed form of the copolymers of benzyl L -aspartate and ethyl L -aspartate results in a transition to the left-handed helix, the temperature of the transition being dependent on the composition of the copolymer. Also poly(β-n-propyl L -aspartate) undergoes a transition from the right- to the left-handed helix form at 59°C. These results suggest a general pattern of behavior of poly(aspartate esters) and that with suitable conditions of solvent and temperature they may be in either the right- or left-handed helical form.  相似文献   

20.
A method for the introduction of side chains containing isonitrile (isocyanide, functional group) on the backbone of polysaccharides and other hydroxylic polymers was developed. The method was based on (a) ionization of some of the hydroxyl groups on the polymer by treatment with a strong base (tert-butoxide) in a polar aprotic solvent (dimethylsulfoxide), and (b) introduction of side chains containing isonitrile groups by nucleophilic attack of the polymeric alkoxide ions on a low molecular weight isonitrile containing a good leaving group in the omega-position, (1-tosyl-3-isocyanopropane). By this method, the side chains containing the-NC functional groups are attached to the polymeric backbone via stable ether bonds. The isonitrile derivatives of cellulose, linear and cross-linked dextran and cross-linked agarose utilized for the covalent fixation of high and low molecular weight ligands by four-component reactions carried out in aqueous medium, at neutral pH.  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