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 812 毫秒
1.
The 13C-NMR spectra of the synthetic membrane modifying nonadecapeptide Boc-(Aib-l-Ala)5-Gly-Ala-Aib-Pro-Ala- Aib-Aib-Glu(OBz)-Gln-OMe (Aib = α-aminoisobutyric acid), and of synthetic intermediates were used for conformational analysis in solution. The assignments of the 13C-NMR signals of Aib are based on the magnetic nonequivalence (MNE) of the geminal Cβ-signals in asymmetric environment resulting in a shift difference of 0.2–0.5 ppm due to neighbouring chiral residues. More than 4 ppm MNE are observed due to α-helical conformation and about 2.5 ppm for Aib situated in the corners of a rigid β-turn. The Ala-Cα signal is also sensitive to different secondary structures. The Cα signal for C-terminal alanine is found at 49–50 ppm, and for alanine within unordered oligopeptides it absorbs at 50–51 ppm. α-Helical environment shifts the Ala-Cα signal to lower field down to 54 ppm. In methanolic solution the nonadecapeptide shows a α-helical N-terminal region. For the C-terminus beginning with proline-14 no periodically ordered conformation is observed, and we suggest a sequence of β-turns. Furthermore the typical E/Z isomerism of the prolyl-peptide bond can be observed on proline itself and on its neighbour alanine.  相似文献   

2.
选择了含奇数碳的环状共轭三烯配体的环庚三烯三羰基钼和铬,在低温下与芳基锂反应,并用Et3OBF4烷基化,得到组成为H8(CO)2MoC(OC2H5)Ar的四个化合物与一个黑色结晶:C7H8(CO)2Cr(OC2H5)H4CF3-P.用元素分析、光谱分析、X射线晶体结构测定研究了它们的结构.  相似文献   

3.
Two (ONO pincer)ruthenium‐complex‐bound norvalines, Boc?[Ru(pydc)(terpy)]Nva?OMe ( 1 ; Boc=tert‐butyloxycarbonyl, terpy=terpyridyl, Nva=norvaline) and Boc?[Ru(pydc)(tBu‐terpy)]Nva?OMe ( 5 ), were successfully synthesized and their molecular structures and absolute configurations were unequivocally determined by single‐crystal X‐ray diffraction. The robustness of the pincer Ru complexes and norvaline scaffolds against acidic/basic, oxidizing, and high‐temperature conditions enabled us to perform selective transformations of the N‐Boc and C?OMe termini into various functional groups, such as alkyl amide, alkyl urea, and polyether groups, without the loss of the Ru center or enantiomeric purity. The resulting dialkylated Ru‐bound norvaline, n‐C11H23CO?l ‐[Ru(pydc)(terpy)]Nva?NH‐n‐C11H23 (l ‐ 4 ) was found to have excellent self‐assembly properties in organic solvents, thereby affording the corresponding supramolecular gels. Ru‐bound norvaline l ‐ 1 exhibited a higher catalytic activity for the oxidation of alcohols by H2O2 than parent complex [Ru(pydc)(terpy)] ( 11 a ).  相似文献   

4.
The kinetics of decarboxylation of six α‐amino acids (glycine, alanine, aminobutyric acid, valine, leucine, and isoleucine) and β‐aminobutyric acid were studied in aqueous solution at 310–330ˆC and 275 bar over the pH25 range 1.5–8.5 by using an in situ FT‐IR spectroscopy flow reactor. Based on the rate of formation of CO2, the first‐order or pseudo‐first‐order rate constants were obtained along with the Arrhenius parameters. The decarboxylation rates of amino acids follow the order Gly > Leu ≈ Ile ≈ Val > Ala > α‐Aib > β‐Aib. Differences in the concentration between 0.05 and 0.5 m had only a minor effect on the decarboxylation rate. The effect of the position of the amino group on the decarboxylation rate was investigated for α‐, β‐, and γ‐aminobutyric acid and the order was found to be α > β ≫ γ. Although the pH dependence is complex, the decarboxylation rates of α‐amino acids qualitatively have the inverse trend of the racemization rates. © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 35: 602–610, 2003  相似文献   

5.
Benzyl N‐[8‐(4,4‐dimethyl‐5‐oxo‐4,5‐dihydrooxazol‐2‐yl)‐2,5,5,8‐tetra­methyl‐3,6‐dioxo‐4,7‐diazanon‐2‐yl]­carbamate, C24H34N4O6, an oxazol‐5(4H)‐one from N‐α‐benzyloxycarbonyl‐(Aib)4‐OH (Aib = α‐amino­isobutyryl) represents the longest peptide oxazolone so far characterized by X‐ray diffraction. The overall geometry of the oxazolone ring compares well with literature data. The Aib(1) and Aib(2) residues are folded into a type III β‐bend, while the conformation adopted by Aib(3), preceding the oxazolone moiety, is semi‐extended. The disposition of the oxazolone ring relative to the preceding residue is stabilized by C—­H?N and C—H?O intramolecular interactions.  相似文献   

6.
The cyclic 16‐membered pentadepsipeptide cyclo(Tro‐Aib‐Aib‐Aib‐Aib) ( 1 ) was crystallized from MeOH/AcOEt/CH2Cl2, and its structure was established by X‐ray crystallography (Fig. 1). There are two symmetry‐independent molecules with different conformations in the asymmetric unit. Two intramolecular H‐bonds stabilize two β‐turns in each molecule. On the other hand, two of the four Aib residues are forced to assume a nonfavorable nonhelical conformation in each of the symmetry‐independent molecules (Table 1). The conformational study in CDCl3 solution by NMR spectroscopy and molecular dynamics (MD) simulations indicate that the averaged structure (Fig. 3) is almost the same as in the solid state.  相似文献   

7.
Cyclization reactions on hexapeptides containing several α‐aminoisobutyric acid (=2‐amino‐2‐methylpropanoic acid; Aib) residues and the turn‐promoting glycine (Gly) and proline (Pro) residues were investigated. Eight linear hexapeptides were synthesized, and their cyclization was attempted with various coupling reagents. The macrolactamization step proved to be difficult since only three hexapeptides could be cyclized. Two of these latter peptides were the linear precursors of the same cyclic hexapeptide, cyclo(Aib‐Aib‐Phe‐Pro‐Aib‐Gly) ( 1 ). Surprisingly, they gave 1 in almost the same yield. Thus, 1 was obtained in 35% yield upon ring closure at the Phe/Pro site by using DEPBT as the coupling reagent, whereas the cyclization at the Aib/Phe site led to 1 in 28 and 34% yield by using PyAOP and DEPC, respectively (DEPBT=3‐[(diethoxyphosphoryl)oxy]‐1,2,3‐benzotriazin‐4(3H)‐one, PyAOP=(1H‐7‐azabenzotriazol‐1‐yloxy)tripyrrolidin‐1‐ylphosphonium hexafluorophosphate, DEPC=diethyl phosphorocyanidate). Another cyclic hexapeptide, cyclo(Aib‐Aib‐Gly‐Aib‐Pro‐Gly) ( 2 ) was prepared in 34% yield when DEPC was used in the cyclization step. The solid‐state conformation of 1 was established by X‐ray crystallography.  相似文献   

8.
Four new Aib‐containing cyclopentapeptides have been synthesized by cyclization of the corresponding linear pentapeptides using the diethyl phosphorocyanidate (DEPC)/EtN(iPr)2 method. The linear precursors were prepared via the ‘azirine/oxazolone method’, i.e., the Aib units were introduced by the reaction of amino acids or peptide acids with a 2,2‐dimethyl‐2H‐azirin‐3‐amine, followed by selective hydrolysis of the terminal amide function. Most remarkably, cyclo[(Aib)5] exists in CDCl3 solution in a symmetrical conformation, i.e., no intramolecular H‐bonds are detectable.  相似文献   

9.
The first synthesis and characterization of Cr(V) complexes of non-sulfur-containing amino acids are reported. The reduction of Cr(VI) in methanol in the presence of amino acids glycine, alanine, and 2-amino-2-methylpropanoic acid (alpha-aminoisobutyric acid, Aib) yielded several Cr(V) EPR signals. For the reaction involving glycine, the only Cr(V) EPR signals detected were those of the Cr(V)-intermediate methanol complexes, which were also observed in the absence of amino acids. The reaction involving alanine yielded one Cr(V) signal with a g(iso) value of 1.9754 (a(iso) = 4.88 x 10(-4) cm(-1) and A(iso)(53Cr) = 17.89 x 10(-4) cm(-1)). However, a solid product isolated from the reaction solution was EPR silent and was characterized as a dioxo-bridged dimeric species, [Cr(V)2(mu-O)2(O)2(Ala)2(OCH3)2](2-), by multiple-scattering XAFS analysis and electrospray mass spectrometry. The EPR spectrum of the reduction reaction of Cr(VI) in the presence of Aib showed several different Cr(V) signals. Those observed at lower g(iso) values (1.9765, 1.9806) were assigned to Cr(V)-methanol intermediates, while the relatively broad six-line signal at g(iso) = 2.0058 was assigned as being due to a Cr(V) complex with coupling to a single deprotonated amine group of the amino acid. This was confirmed by simplification of the superhyperfine coupling lines from six to three when the deuterated ligand was substituted in the reaction. The reduction of Cr(VI) with excess alanine or Aib ligands resulted in the formation of tris-chelate Cr(III) complexes, which were analytically identical to complexes formed via Cr(III) synthesis methods. The fac-[Cr(Aib)3] complex was characterized by single-crystal X-ray diffraction.  相似文献   

10.
Unexpected cleavage of the macrocylic ring of secondary azacrown ethers when interacting with the Aib8 (Aib = alpha-aminoisobutyric acid) oxazolone indicates the possibility for a new mechanism of peptide racemization due to transformations of the oxazolones formed from the N-derivatives of alpha-amino acids in peptide synthesis.  相似文献   

11.
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   

12.
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   

13.
The uncommon amino-acid 2-methylalanine (α-aminoisobutyric acid, Aib) was investigated by 13C-NMR. The chemical shifts of amino- or carboxy-protected derivatives of Aib and of protected oligopeptides are discussed with respect to neighbouring groups and amino acids. The pH-dependence of the 13C-NMR spectra of Aib, Aib-Ala, Ala-Aib, Aib-Ala-Aib and Aib-Ala-Aib-Ala-Aib was studied. Using these examples, a new and advantageous method is demonstrated for the first time for the evalutions of NMR titration curves, which uses so-called chemical shift diagrams (CS diagrams).  相似文献   

14.
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   

15.
《Supramolecular Science》1996,3(1-3):13-18
Hydrophobic helical peptides having alternating hydrophobic amino acid and Aib in the sequence were synthesized to construct supramolecular systems. Three types of supramolecular systems were constructed by the peptides and the derivatives in different environments. First, the dispersion of TFA · H-(Ala-Aib)8-OBzl in water was studied by dynamic light scattering, which suggests the formation of a vesicular structure with an average diameter of 76 nm. We call the peptide assembly in water “peptosome”. Second, Boc-Ser(Ant)-(Ala-Aib)8-OMe spanned the phospholipid bilayer membrane and formed a helix-bundle structure. The bundle structure was supported by ion-channel formation in the membrane. Third, Boc-(Ala-Aib)8-OMe and Boc-(Leu-Aib)8-OBzl formed a two-dimensional crystal at the air-water interface. Boc-(Ala-Aib)12-OBzl also formed a monolayer in a solid state at the air-water interface, but the helix orientation was perpendicular to the interface, which presents a contrast to the parallel orientation of the former hexadecapeptides.  相似文献   

16.
Single‐crystal X‐ray diffraction studies of two terminally protected tetrapeptides Boc‐Ile‐Aib‐Val‐m‐ABA‐OMe ( I ) and Boc‐Ile‐Aib‐Phe‐m‐ABA‐OMe ( II ) (Aib=α‐aminoisobutyric acid; m‐ABA=meta‐aminobenzoic acid) reveal that they form continuous H‐bonded helices through the association of double‐bend (type III and I) building blocks. NMR Studies support the existence of the double‐bend (type III and I) structures of the peptides in solution also. Field emission scanning electron‐microscopic (FE‐SEM) and high‐resolution transmission electron‐microscopic (HR‐TEM) images of the peptides exhibit amyloid‐like fibrils in the solid state. The Congo red‐stained fibrils of peptide I and II , observed between crossed polarizers, show green‐gold birefringence, a characteristic of amyloid fibrils.  相似文献   

17.
The protected poly‐Aib oligopeptides Z‐(Aib)n‐N(Me)Ph with n=2–6 were prepared according to the ‘azirine/oxazolone method’, i.e., by coupling amino or peptide acids with 2,2,N‐trimethyl‐N‐phenyl‐2H‐azirin‐3‐amine ( 1a ) as an Aib synthon (Scheme 2). Following the same concept, the segments Z‐(Aib)3‐OH ( 9 ) and H‐L ‐Pro‐(Aib)3‐N(Me)Ph ( 20 ) were synthesized, and their subsequent coupling with N,N′‐dicyclohexylcarbodiimide (DCC)/ZnCl2 led to the protected heptapeptide Z‐(Aib)3‐L ‐Pro‐(Aib)3‐N(Me)Ph ( 21 ; Scheme 3). The crystal structures of the poly‐Aib oligopeptide amides were established by X‐ray crystallography confirming the 310‐helical conformation of Aib peptides.  相似文献   

18.
The stability and bioavailability of the biomaterial monomers, bisglycidyl methacrylate (BISGMA), bisphenol F diglycidyl ether (BFDGE), and bisphenol A dimethacrylate (BPADM) were investigated using in-vitro techniques. A reverse-phase high-pressure liquid chromatographic/mass spectrometric (HPLC/MS) system was developed to quantitate each monomer and its primary metabolite. Each monomer (10 × 10−6 M ) was incubated at 37 °C under various conditions. Aliquots (N = 3) were removed at various time intervals and quantitated from a standard curve. The in-vitro transport of each parent monomer and its tetrahydroxy metabolite was measured in a Caco-2 system. BISGMA and BPADM were stable in aqueous solution at pH 1. However, BFDGE, was unstable. Plasma esterase of the rat rapidly hydrolyzed the ester compounds, but human esterase did not have a hydrolytic effect on BISGMA or BPADM. BFDGE disappeared in both rat and human plasma, but no tetrahydroxy metabolite was observed. All three parent compounds were unstable in human- and rat-hepatic fractions producing either tetrahydroxy metabolites or bisphenol A (BPA). The tetrahydroxy metabolites, however, were relatively stable under identical conditions, but BPA disappeared when incubated in hepatic-microsomal fractions. While BPADM metabolism produced BPA, an estrogen disrupter, BISGMA and BFDGE did not appear to produce BPA. These results suggest that the potential toxicity of leached dental monomers is more likely to be a result of the metabolite rather than the parent monomer. From Caco-2 transport studies, BFDGE and its tetrahydroxy metabolite both crossed the Caco-2 membrane at a low rate of transport in 2 h (approximately 3 and 5.2%, respectively). The BISGMA metabolite crossed at approximately 8%, indicative of a moderate rate of transport, and BPA crossed at approximately 10% in 1 h (high rate of transport). The transport of BPADM and BISGMA was unable to be determined due to nonspecific absorption to the acrylic vertical transport well. The transport of BFDGE tetrahydroxy metabolite is of particular interest as BFDGE is likely to be chemically hydrolyzed in stomach acid. It is well known that epoxies react with acids resulting in ring opening, so it is not surprising that BFDGE decomposes at pH 1. As a result, it is necessary to identify the decomposition (hydrolysis) products and test their bioavailability.

Mean (SD) stability of BISGMA (10 × 10−6 M ) and bisphenol A tetrahydroxy metabolite in human- and rat-S9-hepatic fractions at 37 °C for 1 h.  相似文献   


19.
Hydrophobic helical peptide molecules with a lactose unit at the C terminal, Nap-(Ala-Aib)(n)-NHCH(2)CH(2)NH-Lac (Nap, Aib, and Lac represent 2-naphthylacetic acid group, 2-aminoisobutyric acid, and lactobionic acid group, respectively, n=4, 6, 8), were synthesized and their formation of self-assemblies in water was investigated. Nap-(Ala-Aib)(4)-NHCH(2)CH(2)NH-Lac was spontaneously dispersed in water and formed aggregates of 70 nm diameter, shown by dynamic light scattering measurement. Cryo-TEM observation revealed that the aggregates took on a vesicular structure with a single membrane. The membrane is suggested to be composed of helical peptide molecules with an interdigitated antiparallel packing on the basis of circular dichroism and fluorescence measurements. On the other hand, the dodecapeptide formed a fibrous assembly, and the hexadecapeptide could not be dispersed in water. Copyright 2000 Academic Press.  相似文献   

20.
According to the earlier published synthesis of the C‐terminal nonapeptide of Trichovirin I 1B, Z‐Ser(tBu)‐Val‐Aib‐Pro‐Aib‐Leu‐Aib‐Pro‐Leuol ( 5 ), the complete tetradecapeptide Z‐Aib‐Asn(Trt)‐Leu‐Aib‐Pro‐Ser(tBu)‐Val‐Aib‐Pro‐Aib‐Leu‐Aib‐Pro‐Leuol ( 11b ), a protected Trichovirin I 1B, has now been prepared by means of the ‘azirine/oxazolone method’. With the exception of the N‐terminal Aib(1), all Aib residues were introduced by the coupling of the corresponding amino or peptide acids with 2,2‐dimethyl‐2H‐azirine‐3‐(N‐methyl‐N‐phenylamine) ( 1a ) and methyl N‐(2,2‐dimethyl‐2H‐azirin‐3‐yl)‐L ‐prolinate ( 3a ) as the Aib and Aib‐Pro synthons, respectively. Single crystals of two segments, i.e., the N‐terminal hexapeptide Z‐Aib‐Asn(Trt)‐Leu‐Aib‐Pro‐Ser(tBu)‐OMe ( 23 ) and the C‐terminal octapeptide Z‐Val‐Aib‐Pro‐Aib‐Leu‐Aib‐Pro‐Leuol ( 17 ), were obtained and their structures have been established by X‐ray crystallography. Following the same strategy, the C‐terminal nonapeptide of Trichovirin I 4A, Z‐Ala‐Val‐Aib‐Pro‐Aib‐Leu‐Aib‐Pro‐Leuol ( 26 ), was also synthesized and characterized by X‐ray crystallography.  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