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 453 毫秒
1.
I. McEwen 《Biopolymers》1993,33(6):933-942
The cyclic hexapeptide cyclo[-Pro1-Gly2-Glu3(OBzl) -Pro4-Phes5-Leu6-] ( 1 ) was modeled and synthesized to be used for chiral discrimination studies. Total correlated spectroscopy and nuclear Overhauser effect spectroscopy spectra of the cyclic hexapeptide 1 in CDCl3 showed the presence of three stereoisomers: two dominant stereoisomers 1a and 1b that exchanged chemically with each other, and a minor stereoisomer 1c (4%) that exchanged exclusively with the stereoisomer 1b . Of the two dominant stereoisomers, only 1a interacted specifically with t-butyloxycarbonyl (Boc-) and 9-flourenylmethyloxycarbonyl (Fmoc-) amino acids in CDCl3. The interaction site of la when complexing with the derivatized amino acids was the chain segment Phe5-Leu6. The Phe5 NH and Leu6 NH protons are contiguous and solvent exposed. Their nmr signals shifted strongly downfield with the addition of Boc- or Fmoc- amino acids to the peptide solution. Thus, both NH protons hydrogen bond to the amino acids, forming a two-point hydrogen-bonding complex. The peptide stereoisomer 1b did not interact specifically with the Boc- and Fmoc-amino acids because of the lack of two contiguous and solvent-exposed peptidic NH protons that seem to be needed for specific interactions of the cyclic hexapeptide 1 with the Boc- and Fmoc-amino acids. A clear difference in the interaction of 1a with D - and L -enantiomers of BocTrp and Fmoc-Trp was observed with nmr spectroscopy. Docking models and molecular mechanics calculations together with nmr observations showed that the NH proton of the indole ring of the Boc-L -Trp and the Fmoc-L- Trp hydrogen bonded to the Pro1 carbonyl group. In this three-point hydrogen-bonding complex, the indole ring becomes locked underneath the Leu residue. The nmr signals of all the Leu6 protons (except for Leu NH) shifted strongly upfield owing to the shielding effect of the indole aromatic ring currents. The indole NH of the D -enantiomer did not hydrogen bond to the Pro1 carbonyl group because the formation of such a three-point hydrogen-bonding complex was thermodynamically unfavorable. © 1993 John Wiley & Sons, Inc.  相似文献   

2.
Ian Mc Ewen 《Biopolymers》1993,33(4):693-702
The cyclic hexapeptide cyclo[-Pro1-Gly2-Glu3(OBzl)-Pro4-Phe5-Leu6-] ( 1 ; OBzl: benzyl ester) was modeled and synthesized to be used as a chiral site for the separation of enantiomers. Total correlation spectroscopy and nuclear Ovehauser effect spectroscopy spectra of the peptide in CDCl3 showed the presence of three stereoisomers. The two dominant stereoisomers 1a and 1b exchanged chemically with each other, while the minor stereoisomer 1c exchanged exclusively with the stereoisomer 1b . Stereoisomer 1a had two cis proline peptide bonds while stereoisomer 1b had all-trans peptide bonds. The stereoisomer 1c had, for nonstrained peptides, an unusual cis phenylalanine peptide bond while both proline peptide bonds were trans. © 1993 John Wiley & Sons, Inc.  相似文献   

3.
Tetrapeptides, Cbz-Gly-X-Y-Gly-OSt ( 1 – 4 )—as well as cyclic systems, cyclo[NH-(CH2)n-CO-Gly-Ser(OX)-Ser(OX)-Gly] ( 5 and 6 ; n = 4 and 2, X = But or H), have been synthesized in order to compare the CD spectrum of linear and cyclic β-turn models containing either a protected or a free hydroxyl of the serine residue. In extremely dilute cyclohexane solution the linear models Cbz-Gly-Ser-Y-Gly-OSt ( 1 – 3a ) show class B spectra with very strong positive bands, contrary to other members of the series. Based on 200-MHz 1H nuclear overhauser enhancement and Fourier transform ir studies, Cbz-Gly-Ser-Ser(OBut)-Gly-OSt ( 3a ) in dilute chloroform solution assumes a distorted type II β-turn conformation fixed by an extended system of intramolecular H bonds. As evidenced by 1H-nmr and FT-IR experiments, the cyclic model cyclo[NH-(CH2)4-CO-Gly-Ser(OBut)- Ser(OBut)-Gly] ( 5a ) in a 1 : 1 mixture of (CD3)2SO-CDCl3 is also characterized by a type II β-turn encompassing the Ser3(OBut)-Gly4 sequence. In water, a class B pattern was measured for this model, in good agreement with theoretical and experimental studies that show that type II β-turns are generally characterized by class B spectra. In the protected and free OH cyclic models, cyclo[NH-(CH2)2-CO-Gly-Ser(OX)-Ser(OX)-Gly] ( 5b and 6b , X = But or H) distortions of the peptide backbone due to the loss of two CH2 groups result in the appearance of CD spectra characterized by a strong negative band near 200 nm, interpreted as a sign of the lack of β-turn structures in these models. This observation, together with other CD data discussed in this paper, clearly show that the CD of serine-containing β-turn sequences strongly depends on long-range backbone and local side-chain interactions.  相似文献   

4.
The crystal structures of the isovaline (Iva) containing dipeptides, Boc-D -Iva-L -Pro-OBz l and Boc-L -Iva-L -Pro-OBz l, were determined by x-ray diffraction. The diastereomeric peptides were shown to adopt unturned conformations closely similar to each other (?Iva 52°, ψIva 46°, ?Pro–65°, and ψPro 143° for D -Iva-L -Pro sequence and ?Iva 52°, ψIva 44°, ?Pro ?63°, and ψpro 148° for L -Iva-L -Pro sequence). The Pro ring of each peptide was in Cγ-endo conformation. The unusually large ∠CIva-NPro-C values (131° in both peptides) were observed, that was due to steric repulsion between the δ-methylene of Pro and the alkyl side chain of Iva residue. These conformations were essentially the same as that of the corresponding α-aminoisobutyric acid (Aib)-containing peptide Boc-Aib-L -Pro-OBz l. The result has demonstrated that replacement of either one of the two methyl groups of the Aib residue in Boc-Aib-L -Pro-OBz l with an ethyl group does not cause any significant change in the unturned conformation of the dipeptide. © 1993 John Wiley & Sons, Inc.  相似文献   

5.
A novel nickel(II) hexaaza macrocyclic complex, [Ni(LR,R)](ClO4)2 ( 1 ), containing chiral pendant groups was synthesized by an efficient one‐pot template condensation and characterized (LR,R═1,8‐di((R)‐α‐methylnaphthyl)‐1,3,6,8,10,13‐hexaazacyclotetradecane). The crystal structure of compound 1 was determined by single‐crystal X‐ray analysis. The complex was found to have a square‐planar coordination environment for the nickel(II) ion. Open framework [Ni(LR,R)]3[C6H3(COO)3]2 ( 2 ) was constructed from the self‐assembly of compound 1 with deprotonated 1,3,5‐benzenetricarboxylic acid, BTC3?. Chiral discrimination of rac‐1,1′‐bi‐2‐naphthol and rac‐2,2,2‐trifluoro‐1‐(9‐anthryl)ethanol was performed to determine the chiral recognition ability of the chiral complex ( 1 ) and its self‐assembled framework ( 2 ). Binaphthol showed a good chiral discrimination on the framework ( 2 ). The optimum experimental conditions for the chiral discrimination were examined by changing the weight ratio between the macrocyclic complex 1 or self‐assembled framework 2 and racemates. The detailed synthetic procedures, spectroscopic data including single‐crystal X‐ray analysis, and the results of the chiral recognition for the compounds are described. Chirality, 25:54‐58, 2013 © 2012 Wiley Periodicals, Inc.  相似文献   

6.
Combinations of L - and D -proline residues are useful compounds for finding new structures and properties of cyclic peptides. This is demonstrated with one striking example, the cyclic tetrapeptide c(D -Pro-L -Pro-D -Pro-L -Pro). For this molecule composed of strictly alternating D - and L -configurated residues, a highly symmetrical structure is expected, which should be an optically inactive meso-form. Cyclization of the enantiomeric pure linear precursor D -Pro-L -Pro-D -Pro-L -Pro, however, yields a racemic mixture of two enantiomeric cyclotetrapeptides, both with twofold symmetry and a cistranscistrans sequence of the peptide bonds. Remarkably, this formation of a racemate was not caused by racemization, but by cis/trans isomerization of all peptide bonds in the ring. This process may occur in the linear precursor during the ring formation (cyclization of conformers with transcistrans or cistranscis arrangement of the amide bonds) as well as in the enantiomeric pure cyclic tetrapeptide at higher temperature. In the latter case, an all-cis structure should exist as the intermediate, which can form a cistranscistrans sequence in two equivalent ways, leading finally to two enantiomeric cyclotetrapeptides. In the first one, the cis peptide bonds are attributed to the L -residues and the trans peptide bonds to the D -residues; in the second one, the cis bonds belong to the D and the trans bonds to the L -residues. The mixture of these two enantiomers does not crystallize in the racemic form, but in enantiomeric pure separate crystals. The structural properties could be proved by 1H- and 13C-nmr spectroscopy and x-ray analysis. The cis/trans isomerization process was confirmed by optical rotation measurements and CD spectroscopy, as well as DREIDING model studies. Calorimetric measurements in the solid state suggest the existence of the expected all-cis intermediate. The backbone conformation of the 12-membered medium-sized ring shows only slight deviations—up to 6° —from the planarity of the peptide bonds. On the other hand, the four pyrrolidine rings show different types of puckering of the Cγ or the Cβ atoms.  相似文献   

7.
We have in the present study explored the anticancer activity against human Burkitt's lymphoma cells (Ramos) of a series of small linear and cyclic tetrapeptides containing a β2,2‐amino acid with either two 2‐naphthyl‐methylene or two para‐CF3‐benzyl side chains, along with their interaction with the main plasma protein human serum albumin (HSA). The cyclic and more amphipathic tetrapeptides revealed a notably higher anticancer potency against Ramos cells [50% inhibitory concentration (IC50) 11–70 μM] compared to the linear tetrapeptide counterparts (IC50 18.7 to >413 μM). The most potent cyclic tetrapeptide c3 had a 16.5‐fold preference for Ramos cells compared to human red blood cells, whereas the cyclic tetrapeptide c1 both showed low hemolytic activity and displayed the overall highest (2.9‐fold) preference for Ramos cells (IC50 23 μM) compared to healthy human lung fibroblast cells (MRC‐5). Investigating the interaction of selected tetrapeptides and recently reported hexapeptides with HSA revealed that the peptides bind to drug site II of HSA in the 22–28 μM range, disregarding size and overall structure. NMR and in silico molecular docking experiments identified the lipophilic residues as responsible for the interaction, but in vitro studies showed that the anticancer potency of the peptides varied in the presence of HSA and that c3 remained the most potent peptide. Based on our findings, we call for implementing serum albumin binding in development of anticancer peptides, as it may have implications for future administration and systemic distribution of peptide‐based cancer drugs. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
The chiral recognition property of poly[(1→6)-2,5-anhydro-3,4-di-O-alkyl-D-glucitol] ( 1 ) toward racemic RCH (CO2CH3)NH3+ · PF6? ( 2 · HPF6) has been studied using a transport system involving an aqueous source and receiving phases separated by a chloroform phase containing 1 . Transport rates for aromatic guests 2a (R = Ph) and 2b (R = CH2Ph) were faster than those for aliphatic guests, 2c (R = CH(CH3)2) and 2d (R = CH2CH(CH3)2), using the polymer substituted with methyl groups ( 1a ). The enantiomeric excess (e.e.) was 10.9% for 2a as a maximum value and decreased in the order of 2a > 2c > 2b = 2d . When the transport of 2a · HPF6 was carried out using the polymers with 3,4-di-O-methyl ( 1a ), ethyl ( 1b ), allyl ( 1c ), and pentyl ( 1d ) groups, the e.e. was 22.0% for 1d as a maximum value and increased in the order of 1a < 1b < 1c < 1d . The formation of a complex between 1a and 2a · HPF6 was confirmed by 1H and 13C NMR spectral measurements. © 1995 Wiley-Liss, Inc.  相似文献   

9.
To obtain general rules of peptide design using α,β-dehydro-residues, a sequence with two consecutive ΔPhe-residues, Boc-L -Val-ΔPhe–ΔPhe- L -Ala-OCH3, was synthesized by azlactone method in solution phase. The peptide was crystallized from its solution in an acetone/water mixture (70:30) in space group P61 with a=b=14.912(3) Å, c= 25.548(5) Å, V=4912.0(6) Å3. The structure was determined by direct methods and refined by a full matrix least-squares procedure to an R value of 0.079 for 2891 observed [I?3σ(I)] reflections. The backbone torsion angles ?1=?54(1)°, ψ1= 129(1)°, ω1=?177(1)°, ?2 =57(1)°, ψ2=15(1)°, ω2 =?170(1)°, ?3=80(1)°, ψ3 =7(2)°, ω3=?177(1)°, ?4 =?108(1)° and ψT4=?34 (1)° suggest that the peptide adopts a folded conformation with two overlapping β-turns of types II and III′. These turns are stabilized by two intramolecular hydrogen bonds between the CO of the Boc group and the NH of ΔPhe3 and the CO of Val1 and the NH of Ala4. The torsion angles of ΔPhe2 and ΔPhe3 side chains are similar and indicate that the two ΔPhe residues are essentially planar. The folded molecules form head-to- tail intermolecular hydrogen bonds giving rise to continuous helical columns which run parallel to the c-axis. This structure established the formation of two β-turns of types II and III′ respectively for sequences containing two consecutive ΔPhe residues at (i+2) and (i+3) positions with a branched β-carbon residue at one end of the tetrapeptide.  相似文献   

10.
Direct enantiomer separation of hypericin, pseudohypericin, and protohypericin was accomplished by high‐performance liquid chromatography (HPLC) using immobilized polysaccharide‐type chiral stationary phases (CSPs). Enantioselectivities up to 1.30 were obtained in the polar‐organic elution mode whereby for hypericin and pseudohypericin Chiralpak IC [chiral selector being cellulose tris(3,5‐dichlorophenylcarbamate)] and for protohypericin Chiralpak IA (chiral selector being the 3,5‐dimethylphenylcarbamate of amylose) gave favorable results. Enantiomers were distinguished by on‐line electronic circular dichroism detection. Optimized enantioselective chromatographic conditions were the basis for determining stereodynamic parameters of the enantiomer interconversion process of hypericin and pseudohypericin. Rate constants delivered by computational simulation of dynamic HPLC elution profiles (stochastic model, consideration of peak tailing) were used to calculate averaged enantiomerization barriers (ΔG) of 97.6–99.6 kJ/mol for both compounds (investigated temperature range 25–45°C). Complementary variable temperature off‐column (i.e., in solution) racemization experiments delivered ΔG = 97.1–98.0 kJ/mol (27–45°C) for hypericin and ΔG = 98.9–101.4 kJ/mol (25–55°C) for pseudohypericin. An activation enthalpy of ΔH# = 86.0 kJ/mol and an activation entropy of ΔS# = ?37.7 J/(K mol) were calculated from hypericin racemization kinetics in solution, whereas for pseudohypericin these figures amounted to 74.1 kJ/mol and ?82.6 J/(K mol), respectively. Although the natural phenanthroperylene quinone pigments hypericin and pseudohypericin as well as their biological precursor protohypericin are chiral and can be separated by enantioselective HPLC low enantiomerization barriers seem to prevent the occurrence of an excess of one enantiomer under typical physiological conditions—at least as long as stereoselective intermolecular interactions with other chiral entities are absent. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

11.
J L Flippen  I L Karle 《Biopolymers》1976,15(6):1081-1092
Chlamydocin, Iabu-L -Phe-D -Pro-L X, is a naturally occurring cyclic tetrapeptide that exhibits high cytostatic activity. The conformation of the peptide ring, as well as the stereo configuration in the vicinity of the epoxide ring, have been established by a single-crystal X-ray study of dihydrochlamydocin: C28H40N4O6·H2O. It crystallizes in the monoclinic space group P21 with a = 12.616(6) Å, b = 12.355(6) Å, c = 9.442(5) Å, and β = 99.5(1)°. The structure was solved by the symbolic addition procedure for phase determination followed by the tangent formula method of phase refinement. This structure represents the first cyclic tetrapeptide in which all four peptide units have been found in the trans conformation; however, each peptide unit is significantly nonplanar with ω angles deviating by 14–24° from the ideal value of 180°. This molecule contains two intramolecular 3 → 1 hydrogen bonds and experimentally determined parameters for these seven-membered turns are presented.  相似文献   

12.
A novel chiral sensor based on the self‐assembled monolayer of (6A‐ω‐mercaptoethylureado‐6A‐deoxy)heptakis(2,3‐di‐o‐phenylcarbamoyl)‐6B, 6C, 6D, 6E, 6F, 6G‐ hexa‐o‐phenylcarbamoyl‐β‐cyclodextrin (Ph‐β‐CD‐SH) on a quartz crystal transducer for chiral recognition was set up. (R,S)‐(±)‐(3‐Methoxyphenyl)ethylamine were recognized by this QCM chiral sensor with a QCM chiral discrimination factor of 1.33. Furthermore, UV spectroscopy was used to investigate the mechanism of host‐guest interactions between (6A‐azido‐6A‐deoxy)heptakis(2,3‐di‐o‐phenylcarbamoyl)‐6B, 6C, 6D, 6E, 6F, 6G‐hexa‐o‐phenylcarbamoyl‐β‐cyclodextrin (Ph‐β‐CD) and (R,S)‐(±)‐(3‐methoxyphenyl) ethylamine. The UV discrimination factor was determined to be 0.066. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

13.
14.
Reaction of melatonin with the hypervalent iron centre of oxoferryl hemoglobin, produced in aqueous solution from methemoglobin and H2O2, has been investigated at 37°C and pH 7.4, by absorption spectroscopy. The reaction results in reduction of the oxoferryl moiety with formation of a heme-ferric containing hemoprotein. Stopped-flow spectrophotometric measurements provide evidence that the reduction of oxoferryl-Hb by melatonin is first-order in oxoferryl-Hb and first-order in melatonin. The bimolecular reaction constant at pH 7.4 and 37°C is 112 ± 1.0 M-1 s-1.

Two major oxidation products from melatonin have been found by gas chromatography-mass spectroscopy: the cyclic compound 1,2,3,3a,8,8a-hexahydro-1-acetyl-5-methoxy-3a-hydroxypyrrolo[2,3-b]indole (cyclic 3-hydroxy-melatonin), and N-acetyl-N′-formyl 5-methoxykynuramine (AFMK). The percentage yield of the two major products appears dependent on the ratio [oxoferryl-Hb]: [melatonin]—the higher the ratio the higher the yield of AFMK. The observed stoichiometry oxoferryl-Hbreduced:melatoninconsumed is 2, when the ratio [oxoferryl-Hb]:[melatonin] is 1:1, but appears >2 at higher molar ratios. The reduction of the hypervalent iron of the oxoferryl moiety may be consistent with an oxidation of melatonin by two one-electron steps.  相似文献   

15.
The crystal structure and conformation of the synthetic cyclic tetrapeptide, cyclo(L -Pro-Sar)2, was determined by x-ray analysis. The peptide crystallizes in the orthorhombic space group P212121 with cell parameters a = 9.277(1), b = 12.884(1), and c = 15.581(2) Å. The crystal structure was solved by the symbolic addition procedure for direct phase determination and least-squares refinement using 1796 reflections, which led to the final R value of 0.043. This structure provides the first example observed in a crystal of a cyclic tetrapeptide in which all four peptide units have been found in the cis conformation with ω angles deviating slightly by 2°–10° from the ideal value of 0°. It was also found that the two Pro Cα-CO single bonds assumed a trans′ (ψ = 159.6° and 158.4°) conformation. Adjoining average planes of the peptide groups fall at nearly right angles to each other. The pyrrolidine ring conformations of the two prolyl residues are in the envelope form, with Cγ carbon out of the least-squares planes for the remaining four atoms.  相似文献   

16.
Triadimefon is a systemic agricultural fungicide of the triazole class whose major metabolite, triadimenol, also a commercial fungicide, provides the majority of the actual fungicidal activity, i.e., inhibition of steroid demethylation. Both chemicals are chiral: triadimefon has one chiral center with two enantiomers while its enzymatic reduction to triadimenol produces a second chiral center and two diastereomers with two enantiomers each. All six stereoisomers of the two fungicides were separated from each other using a chiral BGB‐172 column on a GC‐MS system so as to follow stereospecificity in metabolism by rainbow trout hepatic microsomes. In these microsomes the S‐(+) enantiomer of triadimefon was transformed to triadimenol 27% faster than the R‐(?) enantiomer, forming the four triadimenol stereoisomers at rates different from each other. The most fungi‐toxic stereoisomer (1S,2R) was produced at the slowest rate; it was detectable after 8 h, but below the level of method quantitation. The triadimenol stereoisomer ratio pattern produced by the trout microsomes was very different from that of the commercial triadimenol standard, in which the most rat‐toxic pair of enantiomers (known as “Diastereomer A”) is about 85% of the total stereoisomer composition. The trout microsomes produced only about 4% of “Diastereomer A”. Complementary metabolomic studies with NMR showed that exposure of the separate triadimefon enantiomers and the racemate to rainbow trout for 48 h resulted in different metabolic profiles in the trout liver extracts, i.e., different endogenous metabolite patterns that indicated differences in effects of the two enantiomers. Chirality, 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

17.
Here, we report the general strategies by which NMR spectroscopy can be used to determine the enantiopurity and absolute configuration of chalcogen containing secondary alcohols, including the evaluation of the use of chiral solvating and chiral derivatizing agents. The BINOL/DMAP ternary complex demonstrated a simple and fast protocol for determining enantiopurity. The drug Naproxen afforded a stable, nonhygroscopic, and readily available chiral derivatizing agent (CDA) for NMR chiral discrimination of chalcogen containing secondary alcohols. The chiral recognition by CDA and chiral solvating agent (CSA) was assessed using 1H, 77Se‐{1H}, and 125Te‐{1H} NMR spectroscopy. A simple model for the assignment of the absolute configuration from NMR data is presented.  相似文献   

18.
The natural product cyclic peptide stylissatin A ( 1a ) was reported to inhibit nitric oxide production in LPS‐stimulated murine macrophage RAW 264.7 cells. In the current study, solid‐phase total synthesis of stylissatin A was performed by using a safety‐catch linker and yielded the peptide with a trans‐Phe7‐Pro6 linkage, whereas the natural product is the cis rotamer at this position as evidenced by a marked difference in NMR chemical shifts. In order to preclude the possibility of 1b being an epimer of the natural product, we repeated the synthesis using d ‐allo‐Ile in place of l ‐Ile and a different site for macrocyclization. The resulting product (d ‐allo‐Ile2)‐stylissatin A ( 1c ) was also found to have the trans‐Phe7‐Pro6 peptide conformations like rotamer 1b . Applying the second route to the synthesis of stylissatin A itself, we obtained stylissatin A natural rotamer 1a accompanied by rotamer 1b as the major product. Rotamers 1a , 1b , and the epimer 1c were separable by HPLC, and 1a was found to match the natural product in structure and biological activity. Six related analogs 2–7 of stylissatin A were synthesized on Wang resin and characterized by spectral analysis. The natural product ( 1a ), the rotamer ( 1b ), and (d ‐allo‐Ile2)‐stylissatin A ( 1c ) exhibited significant inhibition of NO.. Further investigations were focused on 1b , which also inhibited proliferation of T‐cells and inflammatory cytokine IL‐2 production. The analogs 2–7 weakly inhibited NO. production, but strongly inhibited IL‐2 cytokine production compared with synthetic peptide 1b . All analogs inhibited the proliferation of T‐cells, with analog 7 having the strongest effect. In the analogs, the Pro6 residue was replaced by Glu/Ala, and the SAR indicates that the nature of this residue plays a role in the biological function of these peptides. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
Conformations of the cyclic tetrapeptide cyclo(L -Pro-Sar)2 in solution were studied by 1H- and 13C-nmr spectrometry and model building. The nmr data provide definite evidence that this cyclic peptide exists chiefly in two conformations, namely, a C2-symmetric conformation and an asymmetric structure. The former was demonstrated to be predominant in polar solvents (100% in Me2SO-d6). This structure contains all cis-peptide bond linkages and all trans′ Pro Cα?CO bonds. It represents the first cyclic tetrapeptide in which all four peptide bonds have been found in the cis-conformation. As the polarity of the solvent decreases, the population of C2-symmetric conformers decreases (88% in CD3CN and 65% in CDCl3). At the same time, a minor asymmetric conformer, characterized by cis-cis-cis-trans peptide bond sequences (two cis Sar-Pro bonds, one cis Pro-Sar bond, and one trans Pro-Sar bond), is seen to increase (9% in CD3CN and 30% in CDCl3). A proposed predominant conformation in solution for cyclo(L -Pro-Sar)2 was compared with a crystal structure, as reported in an accompanying paper. Both structures show striking overall similarities.  相似文献   

20.
Catalytic asymmetric benzylation of a dissymmetric tert‐butylglycinate ketimine, incorporating 1‐naphthyl and phenyl groups as the Schiff base substituents, under phase‐transfer conditions was investigated. It was interesting to note that the sense of asymmetric induction of the alkylation of Z‐imine stereoisomer is opposite to that of the corresponding E stereoisomer with a similar degree of enantioselectivity. More interestingly, the chiral Cu(II) complex of the Schiff base derived from (R)‐2‐phenylglycinol and 2‐hydroxy‐1‐naphthaldehyde was found to catalyze the same reaction under solid‐liquid conditions with comparable enantioselectivity (up to 60% ee) with respect to known cinchona alkaloid catalysts. The solvent/base‐system parameter was shown to control the optimal catalytic activity. Chirality 27:944–950, 2015. © 2015 Wiley Periodicals, Inc.  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