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.
We have described here the self-assembling properties of the synthetic tripeptides Boc-Ala(1)-Aib(2)-Val(3)-OMe 1, Boc-Ala(1)-Aib(2)-Ile(3)-OMe 2 and Boc-Ala(1)-Gly(2)-Val(3)-OMe 3 (Aib=α-amino isobutyric acid, β-Ala=β-alanine) which have distorted β-turn conformations in their respective crystals. These turn-forming tripeptides self-assemble to form supramolecular β-sheet structures through intermolecular hydrogen bonding and other noncovalent interactions. The scanning electron micrographs of these peptides revealed that these compounds form amyloid-like fibrils, the causative factor for many neurodegenerative diseases including Alzheimer's disease, Huntington's disease and Prion-related encephalopathies.  相似文献   

2.
A self-associating synthetic tripeptide [Boc-Ala(1)-Aib(2)-β-Ala(3)-OMe (Aib: α-amino-isobutyric acid, β-Ala: β-alanine)] forms thermoreversible transparent gels in various organic solvents and this offers the first example of a peptide gelator whose molecular self-assembly afforded for gelation has been characterised by single-crystal X-ray diffraction and FT-IR and NMR spectroscopic studies. The crystal structure of an analogous synthetic non-gelator tripeptide [Boc-Ala(1)-Gly(2)-β-Ala(3)-OMe] is also discussed in light of the self-assembly of the gelator tripeptide.  相似文献   

3.
The model tripeptide Boc-beta-Ala(1)-Aib(2)-beta-Ala(3)-OMe 1 [beta-Ala: beta-alanine, Aib: alpha-aminoisobutyric acid] forms an infinite parallel beta-sheet structure through intermolecular interactions in single crystals and from the SEM diagram of this peptide, it is evident that the compound has fibrillar morphology, a characteristic of neurodegenerative disease causing amyloid aggregate.  相似文献   

4.
A relationship between intramolecular hydrogen bonding and the cis-trans isomerization of a proline imide bond for proline-containing short peptides were studied by proton NMR and infrared spectroscopy using DMSO-d6/CDCl3 mixed solvents. The percentage of the trans form increases with increasing fraction of CDCl3 in the mixed solvents except for compounds without possibility of intramolecular hydrogen bonding. Chemical shift variations of amide protons with solvent mixing ratios were found to be useful for judging whether the amide protons take part in the intramolecular hydrogen bonding to a considerable degree or not. These results and infrared spectra were used to specify intramolecularly hydrogen bonded structures of the peptides. Formation of the 10-membered or 13-membered hydrogen bonded ring which includes the carbonyl group precedent to the prolyl residue facilitates the cis-to-trans isomerization and these hydrogen bonded rings are strong enough to restrict the proline imide bond to the trans form in CDCl3 solution. On the other hand, a 7-membered hydrogen bonded ring is not so effective in restricting the proline imide bond.  相似文献   

5.
Single crystal X-ray diffraction studies of a terminally blocked tripeptide Boc-Leu(1)-Aib(2)-Leu(3)-OMe 1 demonstrates that it adopts a bend structure without any intramolecular hydrogen bond. Peptide 1 self-assembles to form a supramolecular antiparallel β-sheet structure by various non-covalent interactions including intermolecular hydrogen bonds in the crystal and it exhibits amyloid-like fibrillar morphology in the solid state.  相似文献   

6.
Four types of α,α-disubstituted amino acids {i.e., α-aminoisobutyric acid (Aib), 1-aminocyclopentanecarboxylic acid (Ac(5)c), (3S,4S)-1-amino-(3,4-dimethoxy)cyclopentanecarboxylic acid [(S,S)-Ac(5)c(dOM)] and its enantiomer (R,R)-Ac(5)c(dOM)} were introduced into l-leucine-based hexapeptides and nonapeptides. The dominant conformations of eight peptides: Cbz-(L-Leu-L-Leu-dAA)(2)-OMe [dAA = 1: Aib; 2: Ac(5)c; 3: (S,S)-Ac(5)c(dOM); 4: (R,R)-Ac(5)c(dOM)] and Boc-(L-Leu-L-Leu-dAA)(3)-OMe [dAA = 5: Aib; 6: Ac(5)c; 7: (S,S)-Ac(5)c(dOM); 8: (R,R)-Ac(5)c(dOM)], were investigated by IR, CD spectra and X-ray crystallographic analysis. The CD spectra revealed that Aib hexapeptide 1 and Ac(5)c hexapeptide 2 formed right-handed (P) 3(10)-helices, while Ac(5)c(dOM) hexapeptides 3 and 4 formed a mixture of (P) 3(10)- and α-helices. The Aib nonapeptide 5 formed a (P) 3(10)-helix, the Ac(5)c nonapeptide 6 formed a mixture of (P) 3(10)- and α-helices, and the Ac(5)c(dOM) nonapeptides 7 and 8 formed (P) α-helices. X-Ray crystallographic analysis revealed that the Aib hexapeptide 1 formed a (P) 3(10)-helix, while (S,S)-Ac(5)c(dOM) hexapeptide 3 formed a (P) α-helix. In addition, the Ac(5)c nonapeptide 6 and (R,R)-Ac(5)c(dOM) nonapeptide 8 formed (P) α-helices. The Aib and achiral Ac(5)c residues have the propensity to form 3(10)-helices in short peptides, whereas the chiral Ac(5)c(dOM) residues have a penchant for forming α-helices.  相似文献   

7.
In order to study the differences of the structural properties of Aib-rich peptides in solution and in the crystalline state, molecular dynamics (MD) simulations of the Aib-containing peptide II (pBrBz-(Aib)5-Leu-(Aib)2-OMe) were performed in the crystalline state, starting from two different conformers obtained experimentally by X-ray diffraction. The structural properties as derived from X-ray crystallography (e.g., torsional angles and hydrogen bonds) are well-reproduced in both constant-volume and constant-pressure simulations, although the force-field parameters used result in a too-high density of the crystals. Through comparison with the results from previous MD and nuclear magnetic resonance (NMR) studies of the very similar peptide I (Z-(Aib)s-Leu-(Aib)2-OMe) in dimethylsulfoxide (DMSO) solution, it is found that, in the crystal simulation, the conformational distribution of peptide II is much narrower than that in the solution simulation of peptide. I. This leads to a significant difference in 3 [symbol: see text] (HN, HC alpha) coupling constant values, in agreement with experimental data, whereas the NOE intensities or proton-proton distance bounds appear insensitive to the difference in conformational distribution. For small peptides the differences between their conformational distribution in the crystalline form and in solution may be much larger than for proteins, a fact which should be kept in mind when interpreting molecular properties in the solution state by using X-ray crystallographic data.  相似文献   

8.
Glycine (Gly) is incorporated in roughly half of all known peptaibiotic (nonribosomally biosynthesized antibiotic peptides of fungal origin) sequences and is the residue with the greatest conformational flexibility. The conformational space of Aib (α‐aminoisobutyric acid) is severely restricted by the second methyl group attached to the Cα atom. Most of the crystal structures containing Aib are N‐terminal protected. Deprotection of the N‐ or C‐terminus of peptides may alter the hydrogen‐bonding scheme and/or the structure and may facilitate crystallization. The structure reported here for glycyl‐α‐aminoisobutyrylglycyl‐α‐aminoisobutyric acid tert‐butyl ester, C16H30N4O5, describes the first N‐terminal‐unprotected (Gly‐Aib)n peptide. The achiral peptide could form an intramolecular hydrogen bond between the C=O group of Gly1 and the N—H group of Aib4. This hydrogen bond is found in all tetrapeptides and N‐terminal‐protected tripeptides containing Aib, apart from one exception. In the present work, this hydrogen bond is not observed (N...O = 5.88 Å). Instead, every molecule is hydrogen bonded to six other symmetry‐related molecules with a total of eight hydrogen bonds per molecule. The backbone conformation starts in the right‐handed helical region (and the left‐handed helical region for the inverted molecule) and reverses the screw sense in the last two residues.  相似文献   

9.
Synthesis and conformational studies of peptides containing the E-vinylogous prolines 1 (VPro1) and 2 (VPro2), Boc-Ala-Val-VPro1-Xaa-Leu-OMe (3, Xaa = Gly; 4, Xaa = Phe), Boc-Ala-Val-VPro2-Xaa-Leu-OMe (5, Xaa = Gly; 6, Xaa = Phe), Boc-Leu-Ile-Val-VPro1-Xaa-Leu-OMe (7, Xaa = Gly; 8, Xaa = Phe), and Boc-Leu-Ile-Val-VPro2-Xaa-Leu-OMe (9, Xaa = Gly; 10, Xaa = Phe), were carried out. It has been shown that both VPro1 and VPro2 lead to the formation of 12-membered intramolecularly hydrogen bonded structures very similar to type VI beta-turns with a cis Xaa-VPro amide bond in the major conformers in all the peptides 3-10, resulting in the nucleation of beta-hairpin type structures in these molecules in CDCl(3).  相似文献   

10.
Three terminally protected short peptides Bis[Boc-D-Leu1-Cys2-OMe] 1, Bis[Boc-Leu1-Cys2-OMe] and Bis[Boc-Val1-Cys2-OMe] 3 exhibit amyloid-like fibrillar morphology. Single crystal X-ray diffraction analysis of peptide 1 clearly demonstrates that it adopts an overall extended backbone molecular conformation that self-assembles to form an intermolecular hydrogen-bonded antiparallel supramolecular beta-sheet structure in crystals. Scanning electron microscopic (SEM) images, transmission electron microscopic (TEM) images and Congo red binding studies vividly demonstrate the amyloid-like fibril formation of peptides 1, 2 and 3. However, after reduction of the disulfide bridge of peptides 1, 2 and 3, three newly generated peptides Boc-D-Leu1-Cys2-OMe 4, Boc-Leu1-Cys2-OMe 5 and Boc-Val1-Cys2-OMe 6 are formed and all of them failed to form any kind of fibril under the same conditions, indicating the important role of the disulfide bond in amyloid-like fibrillogenesis in a peptide model system.  相似文献   

11.
The reaction of Re(NC6H4R)Cl3(PPh3)2 (R = H, 4-Cl, 4-OMe) with 1,2-bis(diphenylphosphino)ethane (dppe) is investigated in refluxing ethanol. The reaction produces two major products, Re(NC6H4R)Cl(dppe)(2)2+ (R = H, 1-H; R = Cl, 1-Cl; R = OMe, 1-OMe) and the rhenium(III) species Re(NHC6H4R)Cl(dppe)2+ (R = H, 2-H; R = Cl, 2-Cl). Complexes 1-H (orthorhombic, Pcab, a = 22.3075(10) A, b = 23.1271(10) A, c = 23.3584(10) A, Z = 8), 1-Cl (triclinic, P1, a = 11.9403(6) A, b = 14.6673(8) A, c = 17.2664(9) A, alpha = 92.019(1) degrees, beta = 97.379(1) degrees, gamma = 90.134(1) degrees, Z = 2), and 1-OMe (triclinic, P1, a = 11.340(3) A, b = 13.134(4) A, c = 13.3796(25) A, alpha = 102.370(20) degrees, beta = 107.688(17) degrees, gamma = 114.408(20) degrees, Z = 1) are crystallographically characterized and show an average Re-N bond length (1.71 A) typical of imidorhenium(V) complexes. There is a small systematic decrease in the Re-N bond length on going from Cl to H to OMe. Complex 2-Cl (monoclinic, Cc, a = 24.2381(11) A, b = 13.4504(6) A, c = 17.466(8) A, beta = 97.06900(0) degrees, Z = 4) is also crystallographically characterized and shows a Re-N bond length (1.98 A) suggestive of amidorhenium(III). The rhenium(III) complexes exhibit unusual proton NMR spectra where all of the resonances are found at expected locations except those for the amido protons, which are at 37.8 ppm for 2-Cl and 37.3 ppm for 1-H. The phosphorus resonances are also unremarkable, but the 13C spectrum of 2-Cl shows a significantly shifted resonance at 177.3 ppm, which is assigned to the ipso carbon of the phenylamido ligand. The extraordinary shifts of the amido hydrogen and ipso carbon are attributed to second-order magnetism that is strongly focused along the axially compressed amido axis. The reducing equivalents for the formation of the Re(III) product are provided by oxidation of the ethanol solvent, which produces acetal and acetaldehyde in amounts as much as 30 equiv based on the quantity of rhenium starting material. Equal amounts of hydrogen gas are produced, suggesting that the catalyzed reaction is the dehydrogenation of ethanol to produce acetaldehyde and hydrogen gas. Metal hydrides are detected in the reaction solution, suggesting a mechanism involving beta-elimination of ethanol at the metal center. Formation of the amidorhenium(III) product possibly arises from migration of a metal hydride in the imidorhenium(V) complex.  相似文献   

12.
Four tripeptides (Z-AA1-2Dpy-AA3-OMe; AA1, AA3 = Gly, Aib) containing a novel amino acid, alpha, alpha-di(2-pyridyl)glycine (2Dpy), were synthesized by the modified Ugi reaction. NMR analysis clearly indicated that the 2Dpy-containing tripeptides except the peptide in which AA1, AA3 = Aib, adopt a unique conformation with two intramolecular hydrogen bonds between 2Dpy-NH and a pyridine nitrogen and between AA3-NH and another pyridine nitrogen. This conformation has so far not been reported. On the other hand, the peptide Z-Aib-2Dpy-Aib-OMe probably adopts a beta-turn structure which is stabilized by two intramolecular hydrogen bonds between 2Dpy-NH and a pyridine nitrogen and between AA3-NH and the C=O of the Z group.  相似文献   

13.
The structural characterization in crystals of three designed decapeptides containing a double d-segment at the C-terminus is described. The crystal structures of the peptides Boc-Leu-Aib-Val-Xxx-Leu-Aib-Val-(D)Ala-(D)Leu-Aib-OMe, (Xxx = Gly 2, (D)Ala 3, Aib 4) have been determined and compared with those reported earlier for peptide 1 (Xxx = Ala) and the all l analogue Boc-Leu-Aib-Val-Ala-Leu-Aib-Val-Ala-Leu-Aib-OMe, which yielded a perfect right-handed alpha-helical structure. Peptides 1 and 2 reveal a right-handed helical segment spanning residues 1 to 7, ending in a Schellman motif with (D)Ala(8) functioning as the terminating residue. Polypeptide chain reversal occurs at residue 9, a novel feature that appears to be the consequence of a C-H.O hydrogen bond between residue 4 C(alpha)H and residue 9 CO groups. The structures of peptides 3 and 4, which lack the pro R hydrogen at the C(alpha) atom of residue 4, are dramatically different. Peptide 3 adopts a right-handed helical conformation over the 1 to 7 segment. Residues 8 and 9 adopt alpha(L) conformations forming a C-terminus type I' beta-turn, corresponding to an incipient left-handed twist of the polypeptide chain. In peptide 4, helix termination occurs at Aib(6), with residues 6 to 9 forming a left-handed helix, resulting in a structure that accommodates direct fusion of two helical segments of opposite twist. Peptides 3 and 4 provide examples of chiral residues occurring in the less favored sense of helical twist; (D)Ala(4) in peptide 3 adopts an alpha(R) conformation, while (L)Val(7) in 4 adopts an alpha(L) conformation. The structural comparison of the decapeptides reported here provides evidence for the role of specific C-H.O hydrogen bonds in stabilizing chain reversals at helix termini, which may be relevant in aligning contiguous helical and strand segments in polypeptide structures.  相似文献   

14.
Single crystal X-ray diffraction studies on the water soluble, synthetic tetrapeptide Tyr(1)-Aib(2)-Tyr(3)-Val(4) with a non-coded amino acid residue (Aib: [small alpha]-amino isobutyric acid) reveal that the peptide adopts an "S"-shaped molecular structure which self-assembles to form a supramolecular triple helix using various non-covalent interactions including water mediated hydrogen bonds in the solid state.  相似文献   

15.
The synthesis of a new conformationally constrained Gly-(s-cis)Pro Turn Mimetic (GPTM) in both racemic and enantiomerically pure forms and their incorporation into peptides 18, 21, and 24 are reported. The synthetic strategy adopted to assemble the bicyclic pyrrolizidinone skeleton is based on the 1,3-dipolar cycloaddition of the cyclic nitrone 4a derived from proline and acrylamide, followed by a reductive cleavage/cyclization domino process. The enantiomerically pure GPTMs are obtained by synthesis and separation of diastereomeric intermediates containing (1R)-1-phenylethylamine as chiral auxiliary. Analysis of pseudotripeptides 18, 21, and 22 by FT-IR and NMR shows that the amide proton of GPTM derivatives 21 is intramolecularly hydrogen bonded in CDCl(3), while DMSO was shown to disrupt this hydrogen bond.  相似文献   

16.
The title compound (systematic name: methyl 2‐{2‐[(tert‐butoxycarbonyl)amino]‐2‐methylpropanamido}‐2‐methylpropanoate), C14H26N2O5, (I), crystallizes in the monoclinic space group P21/n in two polymorphic forms, each with one molecule in the asymmetric unit. The molecular conformation is essentially the same in both polymorphs, with the α‐aminoisobutyric acid (Aib) residues adopting ϕ and ψ values characteristic of α‐helical and mixed 310‐ and α‐helical conformations. The helical handedness of the C‐terminal residue (Aib2) is opposite to that of the N‐terminal residue (Aib1). In contrast to (I), the closely related peptide Boc‐Aib‐Aib‐OBn (Boc is tert‐butoxycarbonyl and Bn is benzyl) adopts an αL‐PII backbone conformation (or the mirror image conformation). Compound (I) forms hydrogen‐bonded parallel β‐sheet‐like tapes, with the carbonyl groups of Aib1 and Aib2 acting as hydrogen‐bond acceptors. This seems to represent an unusual packing for a protected dipeptide containing at least one α,α‐disubstituted residue.  相似文献   

17.
The self-assembly of a terminally protected tripeptide Boc-γ-Abu(1)-Ala(2)-Trp(3)-OMe (γ-Abu = γ-aminobutyric acid) 1 results in the formation of a nanostructured supramolecular zipper through various non-covalent interactions in the crystal in which the indole side-chain of the Trp(3) residue plays a key role via N-H?π interactions.  相似文献   

18.
Recently, novel chiral interactions on 3(10)-helical peptides, of which the helicity is controlled by external chiral stimulus operating on the N-terminus, were proposed as a "noncovalent chiral domino effect (NCDE)" (Inai, Y.; et al. J. Am. Chem. Soc. 2000, 122, 11731. Inai, Y.; et al. J. Am. Chem. Soc. 2002, 124, 2466). The present study clarifies the mechanism for generating the NCDE. For this purpose, achiral nonapeptide (1), H-beta-Ala-(Delta(Z)Phe-Aib)(4)-OMe [Delta(Z)Phe = (Z)-didehydrophenylalanine, Aib = alpha-aminoisobutyric acid], was synthesized. Peptide 1 alone adopts a 3(10)-helical conformation in chloroform. On the basis of the induced CD signals of peptide 1 with chiral additives, chiral acid enabling the predominant formation of a one-handed helix was shown to need at least both carboxyl and urethane groups; that is, Boc-l-amino acid (Boc = tert-butoxycarbonyl) strongly induces a right-handed helix. NMR studies (NH resonance variations, low-temperature measurement, and NOESY) were performed for a CDCl(3) solution of peptide 1 and chiral additive, supporting the view that the N-terminal H-beta-Ala-Delta(Z)Phe-Aib, including the two free amide NH's, captures effectively a Boc-amino acid molecule through three-point interactions. The H-beta-Ala's amino group binds to the carboxyl group to form a salt bridge, while the Aib(3) NH is hydrogen-bonded to either oxygen of the carboxylate group. Subsequently, the free Delta(Z)Phe(2) NH forms a hydrogen bond to the urethane carbonyl oxygen. A semiempirical molecular orbital computation explicitly demonstrated that the dynamic looping complexation is energetically permitted and that the N-terminal segment of a right-handed 3(10)-helix binds more favorably to a Boc-l-amino acid than to the corresponding d-species. In conclusion, the N-terminal segment of a 3(10)-helix, ubiquitous in natural proteins and peptides, possesses the potency of chiral recognition in the backbone itself, furthermore enabling the conversion of the terminally acquired chiral sign and power into a dynamic control of the original helicity and helical stability.  相似文献   

19.
The crystal and molecular structures of the alpha- and beta-L-Asp isomers of L-aspartyl-L-alanine have been determined at 120 K using 1226 and 1609 reflections (I greater than 2.5 sigma I), respectively. The space group for the alpha-isomer is P2(1), with cell parameters a = 4.788(1), b = 16.943(4), c = 5.807(1) A and beta = 107.55(2) degrees; final R factor 0.042. The space group for the beta-isomer is P2(1)2(1)2(1) with a = 4.845(1), b = 9.409(2) and c = 19.170(3) A; final R-factor 0.047. The two peptides crystallize as zwitterions with the side-chain acidic groups ionized. Each molecule adopts a trans configuration at the peptide bond with both carboxyl groups situated on the same side of the peptide plane. The geometries of the aspartyl moieties do, however, differ in the two structures. The peptide bond is significantly longer in the beta-isomer than in the alpha-isomer, with C-N 1.344(3) and 1.328(4) A, respectively. A very short intermolecular carboxyl...carboxyl hydrogen bond (O...O = 2.502(4) A) is observed in the crystals of the alpha-isomer.  相似文献   

20.
Short peptides corresponding to two to four alpha-helical turns of proteins are not thermodynamically stable helices in water. Unstructured octapeptide Ac-His1-Ala2-Ala3-His4-His5-Glu6-Leu7-His8-NH(2) (1) reacts with two [Pd((15)NH(2)(CH(2))(2)(15)NH(2))(NO(3))(2)] in water to form a kinetically stable intermediate, [[Pden](2)[(1,4)(5,8)-peptide]](2), in which two 19-membered metallocyclic rings stabilize two peptide turns. Slow subsequent folding to a thermodynamically more stable two-turn alpha-helix drives the equilibrium to [[Pden](2)[(1,5)(4,8)-peptide]] (3), featuring two 22-membered rings. This transformation from unstructured peptide via turns to an alpha-helix suggests that metal clips might be useful probes for investigating peptide folding.  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