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.
An ir-absorption and Raman-scattering study, in the solid state, has been carried out on monodispersed, N- and C-protected homooligopeptides (number of residues, n, from 2 to 7) of L -valine, L -isoleucine, and L -phenylalanine. The amide I, II, III, V, and vNH regions have been examined. Some deuterated (ND) samples have been examined to complete the assignments. L -Phenylalanine dipeptide displays spectral characteristics compatible with the parallel β-structure; L -isoleucine and L -valine dipeptides are probably in a distorted structure. A mixture of parallel and antiparallel extended chains cannot be excluded for the peptides with n = 3. In the amide I region the spectra of peptides with n ≥ 4 show the existence of the β-conformation. The problem of chain orientation within the pleated-sheet structure is discussed on the basis of a recent theoretical treatment of vibrational interactions of the amide I mode.  相似文献   

2.
The rotational strengths and the robustness values of amide‐I and amide‐II vibrational modes of For(AA)nNHMe (where AA is Val, Asn, Asp, or Cys, n = 1–5 for Val and Asn; n = 1 for Asp and Cys) model peptides with α‐helix and β‐sheet backbone conformations were computed by density functional methods. The robustness results verify empirical rules drawn from experiments and from computed rotational strengths linking amide‐I and amide‐II patterns in the vibrational circular dichroism (VCD) spectra of peptides with their backbone structures. For peptides with at least three residues (n ≥ 3) these characteristic patterns from coupled amide vibrational modes have robust signatures. For shorter peptide models many vibrational modes are nonrobust, and the robust modes can be dependent on the residues or on their side chain conformations in addition to backbone conformations. These robust VCD bands, however, provide information for the detailed structural analysis of these smaller systems. Chirality 27:625–634, 2015 © 2015 Wiley Periodicals, Inc.  相似文献   

3.
M Goodman  C Toniolo  E Peggion 《Biopolymers》1968,6(12):1691-1695
Recent improvements in apparatus permit the examination of circular dichroism (CD) and optical rotatory dispersion (ORD) spectra to 185 mμ. In addition, new solvents which are transparent to 185 mμ have become available for synthetic polypeptides. The spectral region 185–250 mμ is extremely important for the amide (peptide) chromophore, because of the presence at these wavelengths of the n–π* and π–π* bands,1 and of another transition, the assignment of which remains unsettled.2  相似文献   

4.
Vibrational CD (VCD) and ir absorption data are reported for a series of films of Boc-(L -Ala)n-OMe homo-oligopeptides (n = 3–7) in the amide I and A regions. The data evidenced a sharp change between n = 3 and n = 4, which parallels the onset of β-structure formation, and another between n = 5 and n = 6, which parallels the full development of β-structure. This represents the first report of the application of VCD to oligopeptide conformation. The data resembled earlier reported film VCD studies of higher-molecular-weight polypeptides of known β-structure.  相似文献   

5.
The characterization of recently synthesized (Pro-Pro-Gly)n, n = 7, 8 is described, along with melting profile studies of its association equilibrium, and thermal quenching studies of the kinetics of its association reaction. The order of the kinetic reaction is about 3, implying that three peptide chains are involved in the activated state of the rate-limiting step. The reaction rate was found to exhibit a negative temperature coefficient. With the (Pro-Pro-Gly)7 peptide, the concentration dependence of the (Pro-Pro-Gly)n association equilibrium was observed for the first time. Detailed thermodynamic analysis for these n = 7, 8 data, together with literature data for n = 10, 15, 20 were carried out for both the simple “all-or-none” binding model and for a series of complex equilibrium models. For the latter, all of the (Pro-Pro-Gly)n data (in 10% acetic acid) are fit best with a maximally cooperative near-neighbor model with a standard enthalpy change ΔH = ?650 cal/mole of residues, and a standard entropy change ΔS = ?14.63 ?10/n cal/deg-mole of residues, wherein the ?10 eu represents an end-effect contribution to the binding free energy. With regard to optical rotatory properties and thermodynamic parameters, the data for the new n = 7, 8 peptides match rather well with the literature data for the n = 10, 15, and 20 peptides. The enthalpic stabilization per residue of the triple-helical form of (Pro-Pro-Gly)n was nearly an order of magnitude smaller than the enthalpic stabilization per additional proline obtained from direct calorimetric measurements on native collagens of different (and much lower) proline contents by Privalov and Tiktopulo. [Biopolymers (1970) 9 , 127–139.] Possible explanations for this phenomenon are discussed.  相似文献   

6.
Cyclic hexapeptides of the type cyclo(L -Xxx-L -Pro-D -Yyy)2 or cyclo(L -Xxx-L -Pro-Gly)2 exist in solution predominantly in two forms of C2 average symmetry, one with all-trans peptide bonds and generally well-established conformation, and another with both Xxx-Pro peptide bonds cis. We have been measuring the thermodynamic parameters of this equilibrium using carbon and proton nmr spectroscopy. Data have been obtained for peptides in which Yyy = Gly, D -Ala, or D -Phe, and Xxx = Gly, L -Ala, L -Leu, and L -Val. In a given solvent, stability of the all-trans form decreases (ΔG0 increases) as Xxx is changed through the series Gly, L -Ala-, L -Leu, and L -Val, consistent with expected increasing repulsion between the Xxx side chain and the proline δ methylene across the trnas Xxx-Pro bond. Also, for a given set of side chains, the stability of the all-trnas form increases as the polarity of the solvent decreases, consistent with models in which all C?O and N? H groups are accessible for solvation in the two-cis form, but two C?O and two N? H groups are somewhat sequestered in the all-trans form. With the available data it is not possible to identify pure intramolecular (solvent-independent) or pure peptide-bond solvation (side chain-independent) terms in ΔH° or ΔS°, although trends are discernible.  相似文献   

7.
To investigate the structural role played by isostructural unbranched alkyl‐chains on the conformational ensemble and stability of β‐turn structures, the conformational properties of a designed model peptide: Plm‐Pro‐Gly‐Pda ( 1 , Plm: H3C—(CH2)14—CONH—; Pda: —CONH— (CH2)14—CH3) have been examined and compared with the parent peptide: Boc‐Pro‐Gly‐NHMe ( 2 , Boc: tert‐butoxycarbonyl; NHMe: N‐methylamide). The characteristic 13C NMR chemical‐shifts of the Pro Cβ and Cγ resonances ascertained the incidence of an all‐trans peptide‐bond in low polarity deuterochloroform solution. Using FTIR and 1H NMR spectroscopy, we establish that apolar alkyl‐chains flanking a β‐turn promoting Pro‐Gly sequence impart definite incremental stability to the well‐defined hydrogen‐bonded structure. The assessment of 1H NMR derived thermodynamic parameters of the hydrogen‐bonded amide‐NHs via variable temperature indicate that much weaker hydrophobic interactions do contribute to the stability of folded reverse turn structures. The far‐UV CD spectral patterns of 1 and 2 in 2,2,2‐trifluoroethanol are consistent with Pro‐Gly specific type II β‐turn structure, concomitantly substantiate that the flanking alkyl‐chains induce substantial bias in enhanced β‐turn populations. In view of structural as well as functional importance of the Pro‐Gly mediated secondary structures, besides biochemical and biological significance of proteins lipidation via myristoylation or palmytoilation, we highlight potential convenience of the unbranched Plm and Pda moieities not only as main‐chain N‐ and C‐terminal protecting groups but also to mimic and stabilize specific isolated secondary and supersecondary structural components frequently observed in proteins and polypeptides. © 2013 Wiley Periodicals, Inc. Biopolymers 99: 419–426, 2013.  相似文献   

8.
The effect of organic solvents on carboxypeptidase Y (a serine carboxypeptidase from yeast)-catalyzed hydrolysis of amino acid ester and peptide synthesis from N-acyl amino acid ester and amino acid amide was investigated.

The Km value of ester hydrolysis increased with an increase in the solvent content. Dioxane was the most effective and dimethyl sulfoxide (DMSO) the least, whilst Kcat showed a tendency to increase slightly in N, N-dimethylformamide (DMF) and DMSO. For dioxane and acetonitrile (MeCN) a maximum was observed.

In peptide formation from Fua-Phe-OEt and Gly-NH2, dioxane and MeCN supported high product yield at molar fractions smaller than ca. 0.05 but the yield decreased significantly at higher fractions, although a relatively constant selectivity (ratio of the peptide bond formed to the ester consumed) was maintained. DMSO gave rather low peptide yields and selectivity even at lower molar fractions. DMF showed an intermediate tendency.

An apparent saturation parameter of the amine component was evaluated and the dissociation constant of a complex between acyl-enzyme and amino acid amide (Kn), as well as the rate constant of aminolysis exerted by the amino acid amide bound correctly on the enzyme (Kn), was calculated by initial rate analysis of peptide formation. In contrast to Km values, Kn decreased with increasing concentrations of organic cosolvent. while a suppressive effect was observed (except for DMSO) on the Kn parameter.

Effects of the solvent practically immiscible in water was also studied by use of the enzyme physically “immobilized” on glass beads.  相似文献   

9.
M Jackson  H H Mantsch 《Biopolymers》1991,31(10):1205-1212
The structure of valinomycin in a range of organic solvents of varying polarity and in detergent and lipid dispersions has been studied by Fourier transform ir spectroscopy. In solvents of low polarity such as chloroform, ir spectra of valinomycin are fully consistent with the bracelet structure proposed on the basis of nmr spectroscopy, showing a single narrow amide I component attributable to the presence of beta-turns and a single band arising from nonhydrogen-bonded ester C = O groups. K+ complexation results in a downward shift in the amide I band frequency, indicating an increase in the strength of the amide hydrogen bonds, along with a shift to lower frequencies of the ester C = O absorption due to a reduction in electron density in these bonds upon complexation. Identical results were obtained with NH4+, a finding not previously reported. In solvents of both medium (CHCl3/DMSO 3:1) and high (pure DMSO) polarity, we find evidence of significant disruption of the internal hydrogen-bonding network of the peptide and the appearance of a band suggesting the presence of free amide C = O groups. In such solvents, complexation with K+ and NH4+ was not observed. The structure of valinomycin in detergent micelles resembles that in nonpolar organic solvents. However, changes were found in the amide I and ester carbonyl maxima as 2H2O penetrated the micelle which suggest significant interaction between the solvent and peptide. Complexation with K+ was reduced in cationic detergent micelles as a result of a decrease in the effective K+ concentration due to charge repulsion at the micelle surface.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The conventionally protected oligopeptides of the two homologous series Boc-(L -Ile)n-OMe and Boc-(D -aIle)n-OMe (n = 2–6) were synthesized in a standard stepwise fashion and their uv and CD spectra in 2,2,2-trifluoroethanol, and solid-state ir spectra were investigated. In addition, two oligomeric products derived from the NCAs of L -isoleucine and of D -allo-isoleucine and having a DP of 20 and 12, respectively, were studied in the solid state by x-ray and ir. No substantial differences between the properties of the diastereomeric oligomers in the solid state were noticed, a β-structure being very likely at least for the Boc-protected hexapeptides and the higher oligomers. In contrast, differences were observed between the spectroscopic properties of the diastereomeric oligopeptides, and especially of the hexapeptides, in trifluoroethanol solution. The different properties of the hexapeptides in solution were related to the existence, in the case of Boc-(L -Ile)6-OMe, of soluble molecular aggregates in which the peptide chains assume the β-conformation. These results provide an additional example of the influence of the configuration of asymmetric carbon atoms of the side chains on the conformational properties of peptide molecules in solution.  相似文献   

11.
Z-Dehydrophenylalanine (ΔzPhe) possessing four oligopeptides, Boc-(L -Ala-ΔzPhe-Aib)n-OCH3 (n = 1–4: Boc, t-butoxycarbonyl; Aib, α-aminoisobutyric acid), were synthesized, and their solution conformations were investigated by 1H-nmr, ir, uv, and CD spectroscopy and theoretical CD calculation. 1H-nmr (the solvent accessibility of NH groups) and ir studies indicated that all the NH groups except for those belonging to the N-terminal L -Ala-ΔzPhe moiety participate in intramolecular hydrogen bonding in chloroform. This suggests that the peptides n = 2–4 have a 4 → 1 hydrogen-bonding pattern characteristic of 310-helical structures. The uv spectra of all these peptides recorded in chloroform and in trimethyl phosphate showed an intense maximum around 276 nm assigned to the ΔzPhe chromophores. The corresponding CD spectra of the peptides n = 2–4 showed exciton couplets with a negative peak at longer wavelengths, whereas that of the peptide n = 1 showed only weak signals. Theoretical CD spectra were calculated for the peptides n = 2–4 of several helical conformations, on the basis of exciton chirality method. This calculation indicated that the three peptides form a helical conformation deviating from the perfect 310-helix that contains three residues per turn, and that their side chains of Δz Phe residues are arranged regularly along the helix. The center-to-center distance between the nearest phenyl pair(s) was estimated to be ~ 5.5 Å. The chemical shifts of the ΔzPhe side-chain protons (Hβ and aromatic H) for the peptides n = 2–4 indicated anisotropic shielding effect of neighboring phenyl group(s); the effect also supports a regular arrangement of the Δz Phe side chains along the helical axis. © 1993 John Wiley & Sons, Inc.  相似文献   

12.
Self-association of Nα-protected peptides related to C-terminal sequences of substance P in methylene chloride was disrupted by adding increasing amounts of various polar organic solvents. This process was monitored by the disappearance of the amide I C?O stretching band (1630 cm?1) of strongly intermolecularly H-bonded molecules in the irabsorption spectra. The effects induced by main-chain length, incorporation at position 9 of a residue promoting folding (α-aminoisobutyric acid), the nature of solvent, and peptide concentration were established. A corollary 1H-nmr investigation provided detailed information on the NH protons involved in the self-association process as H-bonding donors. The increasing propensity to aggregate exhibited by these peptides is paralleled by a decrease in their solubility. The impact of these results on the synthesis of substance P short sequences is briefly outlined.  相似文献   

13.
The crystal structures of L -3,4-dehydroproline, t-butoxycarbonyl-L -3,4-dehydroproline amide, and acetyl-L -3,4-dehydroproline amide have been determined. L -3,4-Dehydroproline is orthorhombic with a = 16.756, b = 5.870, c = 5.275 Å, and Z = 4; t-butoxycarbonyl-L -3,4-dehydroproline amide is orthorhombic with a = 6.448, b = 8.602, c = 21.710 Å, and Z = 4; acetyl-L -3,4-dehydroproline amide is monoclinic with a = 4.788, b = 10.880, c = 7.785 Å, β = 105.25°, and Z = 2. The final R value for the L -3,4-dehydroproline is 0.046 based on 529 reflections; for t-butoxycarbonyl-L -3,4-dehydroproline amide, 0.050 based on 792 reflections; and for acetyl-L -3,4-dehydroproline amide, 0.058 based on 632 reflections. The structures clearly establish that the free amino acid exists in the zwitterionic form in the crystalline state. The molecular conformations of the t-Boc and acetyl derivatives consist of two planes: one involving the primary amide and the other the remaining atoms of the molecule. The acetyl-L -3,4-dehydroproline amide contains a tertiary amide bond in the cis conformation. To the best of our knowledge, this is the first example of a cis bond in an acetyl derivative of an amino acid or peptide. At variance with the previously reported proline amides, which present ? and ψ values corresponding to those of a right-handed α-helical conformation (conformation A), the t-Boc and acetyl derivatives both have ? and ψ values corresponding to a collagenlike conformation (conformation F).  相似文献   

14.
Vibrational circular dichroism (VCD) spectroscopy was applied to gelation by a chiral low‐molecular mass weight gelator, N,N’‐diperfluoroalkanoyl‐1,2‐trans‐diaminocyclohexane. Attention was focused on the winding effects of (–CF2)n chains on the gelating ability. For this purpose, a series of gelators were synthesized with perfluoroalkyl chains of different length (n = 6–8). When gelation was studied using acetonitrile as a solvent, the fibrils took different morphologies, depending on the chain length: twisted saddle‐like ribbon or helical ribbon from fibril (n = 6) and a helical ribbon from platelet (n = 8). The signs of VCD peaks assigned to the couplet of C=O stretching and to the C‐F stretching were also dependent on n, indicating that a gelator molecule changed conformation on elongating perfluoroalkyl chains. A model is proposed for the aggregation modes in fibrils. Chirality 28:361–364, 2016. © 2016 Wiley Periodicals, Inc.  相似文献   

15.
The dissolution process of model insoluble peptide sequences was investigated in view of the electron acceptor (AN) and electron donor (DN) solvent properties. The Alzheimer's disease-inducing (1-42) Abeta-amyloid peptide and its (1-21) fragment, the (66-97) transmembrane bradykinin B2 receptor sequence, and the strongly aggregated VVLGAAIV were selected as models of insoluble peptides. Solvents presenting similar AN and DN values failed, despite their polarities, to dissociate peptide chains (free in solution or bound to a polymer). The maximum solubility of these aggregated sequences was attained in solvents presenting the highest possible (AN-DN) values (in positive or negative mode). The AN-DN values ranged from approximately -20 to +80 and, notably, the lowest dissociation power was ascribed to solvents presenting values of approximately +40. The strong hydrogen bond donor water is located in this region, indicating that, for dissociation of specific insoluble segments, the solvent should appropriately combine its acid/base strength with the potential for van der Waals interactions. We also observed a sequence-dependent pH effect on peptide solubility confirmed through circular dichroism spectroscopy. This approach also revealed a complex but, in many cases, consistent influence of peptide conformation on its solubility degree, even when structure-inducing solvents were added. In conclusion, the random method of selecting solvents to dissolve insoluble and intractable peptide sequences, still in use by some, could be partially supplanted by the strategy described herein, which may be also applicable to other solute dissociation processes.  相似文献   

16.
We have carried out a systematic analysis of the solid-state conformational preferences of a number of linear homo-oligoprolines (to the tetramer) by ir absorption and x-ray diffraction. The peptides present different chiral sequences (tacticities), various types (urethane and amide) of N-protecting groups, and free and blocked C-termini (which imply different capabilities of forming H-bonds). The following conclusions can be drawn: (i) values for the geometry of the prolyl residue and the peptide bond in the cis and in the trans conformations are proposed; (ii) in general the conformational angles φ and ψ in the linear homo-oligoprolines have values appropriate for the polyproline II structure (conformation F); (iii) the pyrrolidine ring shows various types of puckering with no apparent relation to the backbone conformation; (iv) Pro-Pro peptide bonds generally take the trans conformation, the few cases of cis conformation being formed by Pro residues of different chirality; (v) the single H-bond donor — OH, when present, is always bonded to H-acceptors, which can be either the urethane or the amide or the peptide carbonyl but never the carbonyl group of the — COOH moiety.  相似文献   

17.
The infrared amide I band of collagens (rat and cod skin) and related compounds (polyproline, polyglycine, and polytripeptides) was studied. Assignment of amide I-band components for polyproline II and polytripeptides (Gly-Pro-Pro)n and (Gly-Pro-Gly)n in the solid state and water solution was made. Three amide I components observed in the polypeptide spectra were attributed to three different peptide CO groups in each triplet. On the basis of this assignment, the interpretation of the amide I multicomponent structure in collagen and isomorphous oligo- and polypeptides was attempted. The ordering of intra- and intermolecular hydrogen bonds involving peptide CO groups in collagen and related compounds was discussed.  相似文献   

18.
Original chromosome determinations are presented for 20 American Lupinus taxa, including, for the first time, unifoliolate species, together with first data on meiotic behaviour and pollen fertility for some South American species. Most of the Brazilian multifoliolate L. lanatus, L. rubriflorus, L. multiflorus, L. paranensis, L. bracteolaris and L. reitzii and unifoliolate L. crotalarioides, L. guaraniticus and L. velutinus accessions analysed presented regular chromosome pairing. Meiotic indexes and estimations of pollen viability were higher than 90% for all species and accessions analysed, reflecting the generally regular meiotic behaviour of these plants. Chromosome numbers were determined for the first time for the eastern South‐American species L. guaraniticus, L. crotalarioides, L. paranensis, L. paraguariensis and L. velutinus (n = 18 or 2n = 36) and for the Andean L. ballianus, L. eanophyllus, L. huaronensis, L. semperflorens, plus another eight taxa (2n = 48) from Peru and Bolivia, and L. bandelierae (2n = 36) from Bolivia. Chromosome numbers were confirmed for L. lanatus, L. rubriflorus (2n = 36), L. bracteolaris (2n = 34) and L. microphyllus (2n = 48). In the three accessions of the North American unifoliolate species, L. cumulicola and L. villosus, a chromosome number (2n = 52) previously unknown among American taxa was found. The results of the study, plus published data, support the suggestions that south‐eastern South American species are a group cytologically differentiated from the Andean as well as from most other American ones, and that the Brazilian and the North American unifoliolate Lupinus had independent origins. © 2006 The Linnean Society of London, Botanical Journal of the Linnean Society, 2006, 150 , 229–240.  相似文献   

19.
The preparation of the co-oligopeptides of the series H-Gly-Phe-(Gly)n-Trp-Gly-OH (n = 0, 1, 2) and of other free peptides of glycine, L -tryptophan, and L -phenylalanine is reported. The syntheses have been carried out by conventional methods, using N-hydroxysuccinimide esters for the coupling steps. The ultraviolet absorption properties of the free peptides have been investigated in water. No hypo- or hyperchromicity was found for the aromatic chromophores, with the exception of H-Gly-Phe-Trp-OH, which shows a small but significant hypochromicity. The contribution of the peptide bond to the molar absorptivity in the far ultraviolet has been separated from that of the side chain plus the ? COO? group by plotting the measured molar absorptivity ? of the farthest accessible uv maximum as a function of the number of peptide bonds (nA). The peptide bond contribution proved to be independent of nA in the range nA = 1–5, thus ruling out the onset of helical conformations in the longer chain peptides.  相似文献   

20.
A 13C-nmr study of the salt-induced helix–coil transition of the basic polypeptides poly(L -lysine) [(Lys)n], poly(L -arginine) [(Arg)n], and poly (L -ornithine) [(Orn)n] was performed to serve as a reference of the helical portion of histones and other proteins. As is the case with pH-induced helix–coil transition, the downfield displacement of the Cα and carbonyl carbon signals are observed in the helical state. The upfield shift of the Cβ signals, on the other hand, is noted in the salt-induced transition. Regardless of the differences in the side chains and also the salts used, very similar helix-induced chemical shifts are obtained for (Lys)n and (Arg)n. However, the displacement of the Cα, Cβ, and carbonyl carbons of (Orn)n in the presence of 4M NaClO4 is found to be almost 50% of that of (Lys)n and (Arg)n. This is explained by the fact that the maximum helical content is about 50%, consistent with the ORD result. Further, the motion of the backbone and side chains of the helical from was estimated by measuring the spin-lattice relaxation time (T1), nuclear Overhauser enhancement (NOE), and line width. In the case of (Lys)n, the motion of the side chains is charged very little in comparison with that of the random coil. Indicating that the aggregation of the salt-induced helix is small in contrast to that of the pH-induced helix. For (Arg)n, however, the precipitate of the helical polymers is mainly due to aggregation.  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