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 report herein the first systematic crystal structural investigation of azaproline incorporated in homo- and heterochiral diprolyl peptides. The X-ray crystallography data of peptides 1 – 5 illustrates that stereodynamic nitrogen in azaproline adopted the stereochemistry of neighbouring proline residue without depending on its position in the peptide sequence. Natural bond orbital analysis of crystal structures indicates OazPro−C′Pro of peptides 4 and 5 participating in n→π* interaction with stabilization energy about 1.21–1.33 kcal/mol. Density functional theory calculations suggested that the endo-proline ring puckering favoured over exo-conformation by 6.72–7.64 kcal/mol. NBO and DFT data reveals that the n→π* interactions and proline ring puckering stabilize azaproline chirality with the neighbouring proline stereochemistry. The CD, solvent titration, variable-temperature and 2D NMR experimental results further supported the crystal structures conformation.  相似文献   

2.
Proline derivatives with a C(γ)-exo pucker typically display a high amide bond trans/cis (K(T/C)) ratio. This pucker enhances n→π* overlap of the amide oxygen and ester carbonyl carbon, which favors a trans amide bond. If there were no difference in n→π* interaction between the ring puckers, then the correlation between ring pucker and K(T/C) might be broken. To explore this possibility, proline conformations were constrained using a methylene bridge. We synthesized discrete gauche and anti 5-fluoro- and 5-hydroxy-N-acetylmethanoproline methyl esters from 3-syn and 3-anti fluoro- and hydroxymethanopyrrolidines using directed α-metalation to introduce the α-ester group. NBO calculations reveal minimal n→π* orbital interactions, so contributions from other forces might be of greater importance in determining K(T/C) for the methanoprolines. Consistent with this hypothesis, greater trans amide preferences were found in CDCl(3) for anti isomers en-MetFlp and en-MetHyp (72-78% trans) than for the syn stereoisomers ex-MetFlp and ex-MetHyp (54-67% trans). These, and other, K(T/C) results that we report here indicate how substituents on proline analogues can affect amide preferences by pathways other than ring puckering and n→π* overlap and suggest that caution should be exercised in assigning enhanced pyrrolidine C(γ)-exo ring puckering based solely on enhanced trans amide preference.  相似文献   

3.
The conformations of peptides and proteins are often influenced by glycans O‐linked to serine (Ser) or threonine (Thr). (2S,4R)‐4‐Hydroxyproline (Hyp), together with L ‐proline (Pro), are interesting targets for O‐glycosylation because they have a unique influence on peptide and protein conformation. In previous work we found that glycosylation of Hyp does not affect the N‐terminal amide trans/cis ratios (Ktrans/cis) or the rates of amide isomerization in model amides. The stereoisomer of Hyp—(2S,4S)‐4‐hydroxyproline (hyp)—is rarely found in nature, and has a different influence both on the conformation of the pyrrolidine ring and on Ktrans/cis. Glycans attached to hyp would be expected to be projected from the opposite face of the prolyl side chain relative to Hyp; the impact this would have on Ktrans/cis was unknown. Measurements of 3J coupling constants indicate that the glycan has little impact on the Cγendo conformation produced by hyp. As a result, it was found that the D ‐galactose residue extending from a Cγendo pucker affects both Ktrans/cis and the rate of isomerization, which is not found to occur when it is projected from a Cγexo pucker; this reflects the different environments delineated by the proline side chain. The enthalpic contributions to the stabilization of the trans amide isomer may be due to disruption of intramolecular interactions present in hyp; the change in enthalpy is balanced by a decrease in entropy incurred upon glycosylation. Because the different stereoisomers—Hyp and hyp—project the O‐linked carbohydrates in opposite spatial orientations, these glycosylated amino acids may be useful for understanding of how the projection of a glycan from the peptide or protein backbone exerts its influence.  相似文献   

4.
Recently, the importance of proline ring pucker conformations in collagen has been suggested in the context of hydroxylation of prolines. The previous molecular mechanics parameters for hydroxyproline, however, do not reproduce the correct pucker preference. We have developed a new set of parameters that reproduces the correct pucker preference. Our molecular dynamics simulations of proline and hydroxyproline monomers as well as collagen-like peptides, using the new parameters, support the theory that the role of hydroxylation in collagen is to stabilize the triple helix by adjusting to the right pucker conformation (and thus the right phi angle) in the Y position.  相似文献   

5.
The synthesis, structural properties, and folding patterns of a series of L ‐proline methanologues represented by cis‐ and trans‐4,5‐methano‐L ‐proline amides and their oligomers are reported as revealed by X‐ray crystallography, circular dichroism measurements, and DFT calculations. We disclose the first example of a crystalline tetrameric proline congener to exhibit a polyproline II helical conformation. Experimental evidence of PPII‐type helical arrangement (both in solution and in the solid state) of cis‐4,5‐methano‐L ‐proline oligomers is supported by theoretical calculations reflecting the extent of n→π* stabilization of the trans‐amide conformation.  相似文献   

6.
Cγ-substituted proline derivatives are valuable tools for developing functionalized collagen peptides for biological and materials investigations, yet the stereochemistry at Cγ can produce undesired steric or stereoelectronic constraints. Alkylated γ-azaproline (γ-azPro) derivatives are proline mimetics that lack a stereogenic center at the γ-position of the ring and can thus utilize the invertibility of nitrogen to adapt their conformation. NMR spectroscopic analyses and DFT calculations highlighted how alkylated γ-azPro derivatives are conformationally dynamic and adopt conformational preferences through ring pucker flip along with nitrogen inversion. Lastly, incorporation of alkylated γ-azPro into collagen peptides produced functionalized pH-responsive triple helices with similar thermal stabilities, regardless of their placement in the Xaa or Yaa position within the characteristic Xaa-Yaa-Gly repeating unit of collagen peptides.  相似文献   

7.
Proline residues are critical structural elements in proteins, defining turns, loops, secondary structure boundaries, and polyproline helices. Control of proline conformation therefore may be used to define protein structure and stability. 4-Substituted proline derivatives may be used to control proline ring pucker, which correlates with protein main chain conformation. To examine the use of proline conformational restriction to tune globular protein stability, a series of peptides derived from the trp cage miniprotein was synthesized. Proline at residue 12 of the trp cage miniprotein, which adopts a Cgamma-exo ring pucker in the NMR structure, was replaced with 4-substituted proline derivatives, including 4R derivatives favoring a Cgamma-exo ring pucker and 4S derivatives favoring a Cgamma-endo ring pucker. Eight trp cage peptides were synthesized, five of which included residues that are not commercially available, without requiring any solution phase chemistry. Analysis of the trp cage peptides by circular dichroism and NMR indicated that the structure and stability of the trp cage miniprotein was controllable based on the conformational bias of the proline derivative. Replacement of Pro12 with 4S-substituted proline derivatives that favor the Cgamma-endo ring pucker destabilized the trp cage, while replacement of Pro12 with 4R-substituted proline derivatives that favor a Cgamma-exo ring pucker resulted in increased alpha-helicity and thermal stability of the trp cage. The most stable trp cage derivatives contained benzoates of 4R-hydroxyproline, which also exhibited the most pronounced stereoelectronic effects in TYProxN model peptides. Overall, the stability of the trp cage was tunable by over 50 degrees C depending on the identity of the proline side chain at residue 12.  相似文献   

8.
The title compound, 1‐(2‐deoxy‐β‐d ‐erythro‐pentofuranosyl)‐5‐(prop‐1‐ynyl)pyrimidin‐2,4(1H,3H)‐dione, C12H14N2O5, shows two conformations in the crystalline state: conformer 1 adopts a C2′‐endo (close to 2E; S‐type) sugar pucker and an anti nucleobase orientation [χ = −134.04 (19)°], while conformer 2 shows an S sugar pucker (twisted C2′‐endo–C3′‐exo), which is accompanied by a different anti base orientation [χ = −162.79 (17)°]. Both molecules show a +sc (gauche, gauche) conformation at the exocyclic C4′—C5′ bond and a coplanar orientation of the propynyl group with respect to the pyrimidine ring. The extended structure is a three‐dimensional hydrogen‐bond network involving intermolecular N—H...O and O—H...O hydrogen bonds. Only O atoms function as H‐atom acceptor sites.  相似文献   

9.
Evagelos Bellis 《Tetrahedron》2005,61(36):8669-8676
A series of 4-substituted prolines were prepared and evaluated as organocatalysts for asymmetric aldol reactions. Using (2S,4R)-4-camphorsulfonyloxy-proline, the aldol products were obtained in much higher enantiomeric excess in comparison to that observed using proline itself. In addition, the improved solubility of these new catalysts in organic solvents permits their use in lower sub-stoichiometric amounts compared to proline.  相似文献   

10.
The cyclopropanation of the title compound (S)‐ 2 with various sulfur ylides has been examined. The reaction with methylenesulfonium ylides gave the corresponding cyclopropanes 4 with low diastereoselectivity. The formation of the second product 5 arising from the subsequent methylenation of the CO group was also observed. A clean cyclopropanation of (S)‐ 2 took place with ethyl (dimethylsulfanylidene)acetate affording the cyclopropanes 6 , with high π‐facial selectivity, but low endo/exo ratio. A high endo/exo selectivity, but low π‐facial selectivity was observed in the reaction of (S)‐ 2 with (2‐ethoxy‐2‐oxoethyl)(diphenyl)sulfonium tetrafluoroborate. The use of α‐bromoacetate carbanion as the cyclopropanation reagent resulted in the formation of 6 with very high facial and endo/exo‐selectivity. In a proposed explanation of the stereochemical outcome of the cyclopropanations investigated, the ground‐state conformation of the sulfoxide 2 and the transition‐state structure of the initial addition step were taken into account.  相似文献   

11.
The title compound [systematic name: 4‐amino‐1‐(2‐deoxy‐β‐d ‐erythro‐pentofuranosyl)‐5‐ethynylpyrimidin‐2(1H)‐one], C11H13N3O4, shows two conformations in the crystalline state. The N‐glycosylic bonds of both conformers adopt similar conformations, with χ = −149.2 (1)° for conformer (I‐1) and −151.4 (1)° for conformer (I‐2), both in the anti range. The sugar residue of (I‐1) shows a C2′‐endo envelope conformation (2E, S‐type), with P = 164.7 (1)° and τm = 36.9 (1)°, while (I‐2) shows a major C3′‐exo sugar pucker (C3′‐exo‐C2′‐endo, 3T2, S‐type), with P = 189.2 (1)° and τm = 33.3 (1)°. Both conformers participate in the formation of a layered three‐dimensional crystal structure with a chain‐like arrangement of the conformers. The ethynyl groups do not participate in hydrogen bonding, but are arranged in proximal positions.  相似文献   

12.
The title compound, 4‐amino‐1‐(2‐deoxy‐β‐d ‐erythropentofuranosyl)‐5‐(prop‐1‐ynyl)pyrimidin‐2(1H)‐one, C12H15N3O4, shows two conformations in the crystalline state which differ mainly in the glycosylic bond torsion angle and the sugar pucker. Both mol­ecules exhibit an anti glycosylic bond conformation, with torsion angles χ = −135.0 (2) and −156.4 (2)° for mol­ecules 1 and 2, respectively. The sugar moieties show a twisted C2′‐endo sugar pucker (S‐type), with P = 173.3 and 192.5° for mol­ecules 1 and 2, respectively. The crystal structure is characterized by a three‐dimensional network that is stabilized by several inter­molecular hydrogen bonds between the two conformers.  相似文献   

13.
Photochemistry of Conjugated δ-Keto-enones and β,γ,δ,?-Unsaturated Ketones On 1π,π*-excitation the δ-keto-enones 5–8 are isomerized to compounds B ( 18 , 22 , 26 , 28 ) via 1,3-acyl shift and to compounds C ( 19 , 23 , 27 , 29 ) via 1,2-acyl shift, whereas the β,γ,δ,?-unsaturated ketone 9 gives the isomers 32 and 33 by 1,2-and 1,5-acyl shift, respectively. Furthermore, isomerization of 6 to 24 , dimerization of 8 to 30 and addition of methanol to 8 ( 8 → 31 ) is observed. Unlike 7 and 8 the acyclic ketones 5 , 6 and 9 undergo photodecarbonylation on 1π,π*-excitation ( 5 → 20 , 21 ; 6 → 20 , 25 ; (E)- 9 → 35–38 ). Evidence is given, that the conversion to B as well as the photodecarbonylation of 5,6 and 9 arise from an excited singulet state, but the conversion to C as well as the dimerization of 8 from the T1-state.  相似文献   

14.
A series of novel open-chain and cyclic conformationally constrained α,α-disubstituted (R)- and (S)-glycine derivatives (‘α-chimeras’) combining side chains of Asp, Glu, Leu, Phe, Ser, and Val have been efficiently synthesized by using α-alkylation of racemic 4-monosubstituted 2-phenyl-1,3-oxazol-5(4H)-ones of type 5 , resolution after reaction with (S)-phenylalanine cyclohexylamide ( 8 ) as chiral auxiliary, a novel azlactone/dihydrooxazole interconversion reaction to synthesize optically pure α-substituted (R)- and (S)-serine derivatives coupled with succinimide-ring formation of aspartic-acid derivatives. Based on X-ray structures of (R,S)- 9b , (R,S)- 11c , (R,S)- 18 , and (S,S)- 30 , the absolute configuration of these novel amino-acid building blocks could be unambiguously determined and their preferred conformations in the crystalline state be assessed. The high preference of the open-chain derivatives (R,S)- 1 , (S,S)- 3 , and (R,S)- 11c for β-turn type-I conformations, as well as of the succinimide derivatives (R,S)- 2 , (S,S)- 19 , (S,S)- 24 , (S,S,S)- 26 , and (R,S)- 29 for β-turn type-II conformations and of (S,S)- 4 , (R,S)- 18 , (R,S)- 23 , and (S,S)- 30 for β-turn type-II′ conformations could be confirmed in solution by using CD and NMR spectroscopy. Finally, the spiro derivatives (R,S)- 29 and (S,S)- 30 incorporating the ‘α-chimera’ of Asp/Glu constitute doubly constrained peptide building blocks combining the properties of α-substituted prolines and aspartimides.  相似文献   

15.
A new morpholine-containing foldameric hybrid peptide was synthesized in solution phase, and the conformational preferences were assessed by means of NMR and molecular modeling calculations. All data suggested the existence of two equilibrating conformations involving hydrogen-bonds in the major rotamer. Moreover, calculations on higher model foldamers indicated seven-membered ring hydrogen-bond forming γ-turns as the main driving force in the stabilization of helix-folded conformations. Thus, this study suggests the possibility of using morpholine-3-COOH as a proline surrogate to generate higher α/β hybrid peptides.  相似文献   

16.
When α,β-unsaturated γ-dimethoxymethyl cyclohexenones are excited to the S2(π,π*) state, certain unimolecular reactions can be observed to compete with S2 → S1 internal conversion. These reactions do not occur from the S1(n,π*) or the lowest T(π,π* and n,π*) states. They comprise the radical elimination of the formylacetal substituent (cf. 8 , 9 → 32 + 33 ), γ → α formylacetal migration (cf. 6 → 27 , 8 → 30 , 9 → 34 , 12 → 37 ), and a cyclization process involving the transfer of a methoxyl hydrogen to the α carbon and ring closure at the β position (cf. 6 → 28 , 8 → 31 , 12 → 38 , 20 → 40 + 41 ). The quantum yield of the ring closure 20a → 40a + 41a is 0.016 at ≤ 0.05M concentration. It is independent of the excitation wavelength within the π→π* absorption band (238–254 nm), but Φ ( 40a + 41a ) decreases at higher concentrations. According to the experimental data the reactive species of these specifically π→π*-induced transformations is placed energetically higher than the S1(n,π*) state, and it is either identical with the thermally equilibrated S2(n,π*) state, or reached via this latter state. The linear dienone 14 undergoes a similar π→π*-induced cyclization (→ 42 ) whereas the benzohomologue 26 proved unreactive, and the dienone 22 at both n → π and π→π* excitation only gives rise to rearrangements generally characteristic of cross-conjugated cyclohexadienones.  相似文献   

17.
The two components of the dual phosphorescence of 1-indanone ( 1 ) and six related ketones ( 2–7 ) possess different excitation spectra exhibiting the vibrational progression characteristic of the S0 → S1 (n, π*) transition (shorter-lived emission) and two bands of the S0 → S2 and 3 (π,π*) 0–0 transitions, respectively. The most favorable intersystem crossing routes are S1 (n, π*) → T (n, π*) and S2,3 (π*) → T (π, π*). Internal conversion to S1 competes more effectively with S (π, π*) → T (π, π*) intersystem crossing only from higher vibrational levels of the S2 and S3 states.  相似文献   

18.
Abstract

The decarboxylation reaction of δ -cis-β-[Co(L1)(pdH)]2+ complex yielded δ -cis-β-[Co(L1) (R-pro)]2+, while the δ -cis-β-[Co(L2) (S-pro)]2+ was obtained from the reaction of δ -cis-β-[Co(L2) (pdH)]2+, where L1 is (3R)3-methyl-1, 6-bis[(2S)-pyrrolidin-2-yl]-2, 5-diazahexane, L2 is (3S) 3-methyl-1, 6-bis-[(2S)-pyrrolidin-2-yl]-2, 5-diazahexane, and pdH is the pyrrolidine-2, 2-dicarboxylate ion. The asymmetrically synthesized prolines were isolated via the decomposition of the decarboxylated complexes. The proline isolated from δ -cis-β-[Co(L1) (R-pro)]2+ showed a specific rotation of +12.0, representing a 24% excess of R-proline over S-proline, while the proline isolated from δ -cis-β-[Co(L2) (S-pro)]2+ showed a specific rotation of -10.0, indicating a 20% excess of S-proline over R-proline.  相似文献   

19.
The crystal structures of captopril {systematic name: (2S)‐1‐[(2S)‐2‐methyl‐3‐sulfanylpropanoyl]pyrrolidine‐2‐carboxylic acid}, C9H15NO3S, (1), and its dimer disulfide metabolite, 1,1′‐{disulfanediylbis[(2S)‐2‐methyl‐1‐oxopropane‐3,1‐diyl]}bis‐L‐proline, C18H28N2O6S2, (2), were determined by single‐crystal X‐ray diffraction analysis. Compound (1) crystallizes in the orthorhombic space group P212121, while compound (2) crystallizes in the monoclinic space group P21, both with one molecule per asymmetric unit. The molecular geometries of (1) and (2) are quite similar, but certain differences appear in the conformations of the five‐membered proline rings and the side chains containing the sulfhydryl group. The proline ring adopts an envelope conformation in (1), while in (2) it exists in envelope and slightly deformed half‐chair conformations. The conformation adopted by the side chain is extended in (1) and folded in (2). A minimum‐energy conformational search using Monte Carlo methods in the aqueous phase reveals that the optimized conformations of the title compounds differ from those determined crystallographically, which depend on their immediate environment. Intermolecular O—H...O and relatively weak C—H...O interactions seem to be effective in both structures and, together with S—H...O and C—H...S contacts, they create three‐dimensional networks.  相似文献   

20.
Herein, we report that the conformational preference of phenyl acetate is governed by steric effect and n→π* interaction. Conformation‐specific electronic and IR spectroscopy combined with quantum chemistry calculations confirm the presence of only the cis conformer of phenyl acetate in the experiment. The cis conformer of phenyl acetate has n→π* interaction between the lone‐pair electrons on the carbonyl oxygen atom and the π* orbitals of the phenyl group. The n→π* interaction is absent in the trans conformer which has additional steric repulsion between the methyl group and phenyl ring. The trans conformer is higher in energy than the cis conformer by ≈3 kcal mol?1. We have found the effect of methyl substitution on the strength of the n→π* interaction, steric repulsion, and hyperconjugation in phenyl acetate. The red‐shift observed in the cis conformer of phenyl acetate with respect to the trans conformer is affected due to the influence of the methyl substituent on the strength of the n→π* interaction as well as hyperconjugation. The present result demonstrates that the introduction of a bulkier substituent can induce steric as well as electronic control to reduce conformational heterogeneity of a molecular system. Understanding the effect of bulkier substituents to promote defined conformations having specific non‐covalent interactions may have implication in better perception of the optimum structure and function of biomolecules as well as recognition of drugs by biomolecules.  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