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.
《化学:亚洲杂志》2017,12(24):3195-3202
Herein we report the unique conformations adopted by linear and cyclic tetrapeptides (CTPs) containing 2‐aminobenzoic acid (2‐Abz) in solution and as single crystals. The crystal structure of the linear tetrapeptide H2N‐d ‐Leu‐d ‐Phe‐2‐Abz‐d ‐Ala‐COOH ( 1 ) reveals a novel planar peptidomimetic β‐turn stabilized by three hydrogen bonds and is in agreement with its NMR structure in solution. While CTPs are often synthetically inaccessible or cyclize in poor yield, both 1 and its N ‐Me‐d ‐Phe analogue ( 2 ) adopt pseudo‐cyclic frameworks enabling near quantitative conversion to the corresponding CTPs 3 and 4 . The crystal structure of the N ‐methylated peptide ( 4 ) is the first reported for a CTP containing 2‐Abz and reveals a distinctly planar 13‐membered ring, which is also evident in solution. The N ‐methylation of d ‐Phe results in a peptide bond inversion compared to the conformation of 3 in solution.  相似文献   

2.
The crystal structure of 5‐fluorosalicylic acid is known from the literature [Choudhury & Guru Row (2004). Acta Cryst. E 60 , o1595–o1597] as crystallizing in the monoclinic crystal system with space‐group setting P21/n and with one molecule in the asymmetric unit (polymorph I). We describe here a new polymorph which is again monoclinic but with different unit‐cell parameters (polymorph II). Polymorph II has two molecules in the asymmetric unit. Its structure was modelled as a twin, with a pseudo‐orthorhombic C‐centred twin cell.  相似文献   

3.
Molecular salts and cocrystals of amino acids have potential applications as molecular materials with nonlinear optical, ferroelectric, piezoelectric, and other various target physical properties. The wide choice of amino acids and coformers makes it possible to design various crystal structures. The amino acid–maleic acid system provides a perfect example of a rich variety of crystal structures with different stoichiometries, symmetries and packing motifs built from the molecular building blocks, which are either exactly the same, or differ merely by protonation or as optical isomers. The present paper reports the crystal structures of two new salts of the dl ‐norvaline–maleic acid system with 1:1 and 2:1 stoichiometries, namely dl ‐norvalinium hydrogen maleate, C5H12NO2+·C4H3O4, (I), and dl ‐norvalinium hydrogen maleate–dl ‐norvaline, C5H12NO2+·C4H3O4·C5H11NO2, (II). These are the first examples of molecular salts of dl ‐norvaline with an organic anion. The crystal structure of (I) has the same C 22(12) structure‐forming motif which is common for hydrogen maleates of amino acids. The structure of (II) has dimeric cations. Of special interest is that the single crystals of (I) which are originally formed on crystallization from aqueous solution transform into single crystals of (II) if stored in the mother liquor for several hours.  相似文献   

4.
Recently, amino acid ionic liquids (AAILs) have attracted much research interest. In this paper, we present the first application of AAILs in chiral separation based on the chiral ligand exchange principle. By using 1‐alkyl‐3‐methylimidazolium L ‐proline (L ‐Pro) as a chiral ligand coordinated with copper(II), four pairs of underivatized amino acid enantiomers—dl ‐phenylalanine (dl ‐Phe), dl ‐histidine (dl ‐His), dl ‐tryptophane (dl ‐Trp), and dl ‐tyrosine (dl ‐Tyr)—were successfully separated in two major chiral separation techniques, HPLC and capillary electrophoresis (CE), with higher enantioselectivity than conventionally used amino acid ligands (resolution (Rs)=3.26–10.81 for HPLC; Rs=1.34–4.27 for CE). Interestingly, increasing the alkyl chain length of the AAIL cation remarkably enhanced the enantioselectivity. It was inferred that the alkylmethylimidazolium cations and L ‐Pro form ion pairs on the surface of the stationary phase or on the inner surface of the capillary. The ternary copper complexes with L ‐Pro are consequently attached to the support surface, thus inducing an ion‐exchange type of retention for the dl ‐enantiomers. Therefore, the AAIL cation plays an essential role in the separation. This work demonstrates that AAILs are good alternatives to conventional amino acid ligands for ligand‐exchange‐based chiral separation. It also reveals the tremendous application potential of this new type of task‐specific ILs.  相似文献   

5.
6.
We report a novel 1:1 cocrystal of β‐alanine with dl ‐tartaric acid, C3H7NO2·C4H6O6, (II), and three new molecular salts of dl ‐tartaric acid with β‐alanine {3‐azaniumylpropanoic acid–3‐azaniumylpropanoate dl ‐tartaric acid–dl ‐tartrate, [H(C3H7NO2)2]+·[H(C4H5O6)2], (III)}, γ‐aminobutyric acid [3‐carboxypropanaminium dl ‐tartrate, C4H10NO2+·C4H5O6, (IV)] and dl ‐α‐aminobutyric acid {dl ‐2‐azaniumylbutanoic acid–dl ‐2‐azaniumylbutanoate dl ‐tartaric acid–dl ‐tartrate, [H(C4H9NO2)2]+·[H(C4H5O6)2], (V)}. The crystal structures of binary crystals of dl ‐tartaric acid with glycine, (I), β‐alanine, (II) and (III), GABA, (IV), and dl ‐AABA, (V), have similar molecular packing and crystallographic motifs. The shortest amino acid (i.e. glycine) forms a cocrystal, (I), with dl ‐tartaric acid, whereas the larger amino acids form molecular salts, viz. (IV) and (V). β‐Alanine is the only amino acid capable of forming both a cocrystal [i.e. (II)] and a molecular salt [i.e. (III)] with dl ‐tartaric acid. The cocrystals of glycine and β‐alanine with dl ‐tartaric acid, i.e. (I) and (II), respectively, contain chains of amino acid zwitterions, similar to the structure of pure glycine. In the structures of the molecular salts of amino acids, the amino acid cations form isolated dimers [of β‐alanine in (III), GABA in (IV) and dl ‐AABA in (V)], which are linked by strong O—H…O hydrogen bonds. Moreover, the three crystal structures comprise different types of dimeric cations, i.e. (AA)+ in (III) and (V), and A+A+ in (IV). Molecular salts (IV) and (V) are the first examples of molecular salts of GABA and dl ‐AABA that contain dimers of amino acid cations. The geometry of each investigated amino acid (except dl ‐AABA) correlates with the melting point of its mixed crystal.  相似文献   

7.
Alkanolamines have been known for their high CO2 absorption for over 60 years and are used widely in the natural gas industry for reversible CO2 capture. In an attempt to crystallize a salt of (RS)‐2‐(3‐benzoylphenyl)propionic acid with 2‐amino‐2‐methylpropan‐1‐ol, we obtained instead a polymorph (denoted polymorph II) of bis(1‐hydroxy‐2‐methylpropan‐2‐aminium) carbonate, 2C4H12NO+·CO32−, (I), suggesting that the amine group of the former compound captured CO2 from the atmosphere forming the aminium carbonate salt. This new polymorph was characterized by single‐crystal X‐ray diffraction analysis at low temperature (100 K). The salt crystallizes in the monoclinic system (space group C2/c, Z = 4), while a previously reported form of the same salt (denoted polymorph I) crystallizes in the triclinic system (space group P, Z = 2) [Barzagli et al. (2012). ChemSusChem, 5 , 1724–1731]. The asymmetric unit of polymorph II contains one 1‐hydroxy‐2‐methylpropan‐2‐aminium cation and half a carbonate anion, located on a twofold axis, while the asymmetric unit of polymorph I contains two cations and one anion. These polymorphs exhibit similar structural features in their three‐dimensional packing. Indeed, similar layers of an alternating cation–anion–cation neutral structure are observed in their molecular arrangements. Within each layer, carbonate anions and 1‐hydroxy‐2‐methylpropan‐2‐aminium cations form planes bound to each other through N—H…O and O—H…O hydrogen bonds. In both polymorphs, the layers are linked to each other via van der Waals interactions and C—H…O contacts. In polymorph II, a highly directional C—H…O contact (C—H…O = 156°) shows as a hydrogen‐bonding interaction. Periodic theoretical density functional theory (DFT) calculations indicate that both polymorphs present very similar stabilities.  相似文献   

8.
A new crystalline form of benzene‐1,2‐diamine, C6H8N2, crystallizing in the space group Pbca, has been identified during screening for cocrystals. The crystals are constructed from molecular bilayers parallel to (001) that have the polar amino groups directed to the inside and the aromatic groups, showing a herringbone arrangement, directed to the outside. The known monoclinic form and the new orthorhombic polymorph exhibit two‐dimensional isostructurality as the crystals consist of nearly identical bilayers. In the monoclinic form, neighbouring bilayers are generated by a unit translation along the a axis, whereas in the orthorhombic form they are generated by a c‐glide. Moreover, the new form of benzene‐1,2‐diamine is essentially isomorphous with the only known form of 2‐aminophenol.  相似文献   

9.
Polymorph VI of 4‐amino‐N‐(2‐pyridyl)benzenesulfonamide, C11H11N3O2S, is monoclinic (space group P21/n). The asymmetric unit contains two different tautomeric forms. The structure displays N—H...N and N—H...O hydrogen bonding. The two independent molecules form two separate two‐ and three‐dimensional hydrogen‐bonded networks which interpenetrate. The observed patterns of hydrogen bonding are analogous to those in polymorph I of sulfathiazole.  相似文献   

10.
A new polymorph of bis(2‐aminopyridinium) fumarate–fumaric acid (1/1), 2C5H7N2+·C4H2O42−·C4H4O4, was obtained and its crystal structure determined by powder X‐ray diffraction. The new polymorph (form II) crystallizes in the triclinic system (space group P), while the previous reported polymorph [form I; Ballabh, Trivedi, Dastidar & Suresh (2002). CrystEngComm, 4 , 135–142; Büyükgüngör, Odabaşoğlu, Albayrak & Lönnecke (2004). Acta Cryst. C 60 , o470–o472] is monoclinic (space group P21/c). In both forms I and II, the asymmetric unit consists of one 2‐aminopyridinium cation, half a fumaric acid molecule and half a fumarate dianion. The fumarate dianion is involved in hydrogen bonding with two neighbouring 2‐aminopyridinium cations to form a hydrogen‐bonded trimer in both forms. In form II, the hydrogen‐bonded trimers are interlinked across centres of inversion via pairs of N—H...O hydrogen bonds, whereas such trimers are joined via single N—H...O hydrogen bonds in form I, leading to different packing modes for forms I and II. The results demonstrate the relevance and application of the powder diffraction method in the study of polymorphism of organic molecular materials.  相似文献   

11.
The interaction of [Ru(η6‐C10H8)(Cp)]+ (Cp=C5H5) with aromatic amino acids (L ‐phenylalanine, L ‐tyrosine, L ‐tryptophane, D ‐phenylglycine, and L ‐threo‐3‐phenylserine) under visible‐light irradiation gives the corresponding [Ru(η6‐amino acid)(Cp)]+ complexes in near‐quantitative yield. The reaction proceeds in air at room temperature in water and tolerates the presence of non‐aromatic amino acids (except those which are sulfur containing), monosaccharides, and nucleotides. The complex [Ru(η6‐C10H8)(Cp)]+ was also used for selective labeling of Tyr and Phe residues of small peptides, namely, angiotensin I and II derivatives.  相似文献   

12.
Biological macromolecules are essentially homochiral. For example, proteins mostly consist of l ‐amino acids. What happens when a chiral molecule meets itself in a mirror? For expanded polyvaline, zigzag‐helix transformation occurs. In this study, expanded polyvalines containing bis(pyridine)silver(I) moieties were synthesized and isolated as single crystals. The molecular structures were determined by X‐ray analysis, which revealed that chiral expanded poly(l ‐valine) and poly(d ‐valine) form zigzag chains. However, racemic mixture of these molecules form left‐ and right‐handed 41 helices that retain the original sequences. These secondary structures can be transformed by only flipping the C‐terminal amide plane for each unit, which is reminiscent of the relationship between an α‐helix and a β‐strand. Such expanded polypeptides can be built up into expanded protein, forming a tailor‐made three‐dimensional structure, which will lead to new functions.  相似文献   

13.
The crystal structure of N‐(l ‐2‐amino­butyryl)‐l ‐alanine, C7H14N2O3, is closely related to the structure of l ‐alanyl‐l ‐alanine, both being tetragonal, while the retro‐analogue 2‐(l ‐alanyl­amino)‐l ‐butyric acid 0.33‐hydrate, C7H14N2O3·­0.33H2O, forms a new type of molecular columnar structure with three peptide mol­ecules in the asymmetric unit.  相似文献   

14.
Ganciclovir (GCV; systematic name: 2‐amino‐9‐{[(1,3‐dihydroxypropan‐2‐yl)oxy]methyl}‐6,9‐dihydro‐1H‐purin‐6‐one), C9H13N5O4, an antiviral drug for treating cytomegalovirus infections, has two known polymorphs (Forms I and II), but only the structure of the metastable Form II has been reported [Kawamura & Hirayama (2009). X‐ray Struct. Anal. Online , 25 , 51–52]. We describe a successful preparation of GCV Form I and its crystal structure. GCV is an achiral molecule in the sense that its individual conformers, which are generally chiral objects, undergo fast interconversion in the liquid state and cannot be isolated. In the crystalline state, GCV exists as two inversion‐related conformers in Form I and as a single chiral conformer in Form II. This situation is similar to that observed for glycine, also an achiral molecule, whose α‐polymorph contains two inversion‐related conformers, while the γ‐polymorph contains a single conformer that is chiral. The hydrogen bonds are exclusively intermolecular in Form I, but both inter‐ and intramolecular in Form II, which accounts for the different molecular conformations in the two polymorphs.  相似文献   

15.
Racemic 2,4‐di‐O‐benzoyl‐myo‐inositol‐1,3,5‐orthoacetate, which normally crystallizes in a monoclinic form (form I, space group P21/n) could be persuaded to crystallize out as a metastable polymorph (form II, space group C2/c) by using a small amount of either D ‐ or L ‐ 2,4‐di‐O‐benzoyl‐myo‐inositol‐1,3,5‐orthoformate as an additive in the crystallization medium. The structurally similar enantiomeric additive was chosen by the scrutiny of previous experimental results on the crystallization of racemic 2,4‐di‐O‐benzoyl‐myo‐inositol‐1,3,5‐orthoacetate. Form II crystals can be thermally transformed to form I crystals at about 145 °C. The relative organization of the molecules in these dimorphs vary slightly in terms of the helical assembly of molecules, that is, electrophile (El)???nucleophile (Nu) and C? H???π interactions, but these minor variations have a profound effect on the facility and specificity of benzoyl‐group‐transfer reactivity in the two crystal forms. While form II crystals undergo a clean intermolecular benzoyl‐group‐transfer reaction, form I crystals are less reactive and undergo non‐specific benzoyl‐group transfer leading to a mixture of products. The role played by the additive in fine‐tuning small changes that are required in the molecular packing opens up the possibility of creating new polymorphs that show varied physical and chemical properties. Crystals of D ‐2,6‐di‐O‐benzoyl‐myo‐inositol‐1,3,5‐orthoformate (additive) did not show facile benzoyl‐group‐transfer reactivity (in contrast to the corresponding racemic compound) due to the lack of proper juxtaposition and assembly of molecules.  相似文献   

16.
3,6‐Dinitrodurene (1,2,4,5‐tetramethyl‐3,6‐dinitrobenzene), C10H12N2O4, has been crystallized in two polymorphic forms which may be distinguished by their colours in the solid state. Polymorph I gives clear colourless prismatic crystals, while polymorph II crystallizes in the dark and under an inert atmosphere as irregular purple blocks. Both forms belong to the space group C2/c, with both asymmetric units containing two half‐molecules. One molecule is located on an inversion centre and the other lies on a twofold axis. The polymorphism arises from different orientations of the twofold axis: in form I, this axis passes through the mid‐points of two C—C bonds of the benzene ring and, as a consequence, all atoms in the asymmetric unit are in general positions. In form II, the N atoms of the nitro groups and the Cipso atoms are located on the binary axis. Comparing phases I and II, slightly different conformations are observed for the nitro substituents, while the stacking structures are very similar.  相似文献   

17.
Two new cyclic tetrapeptides, cyclo(l ‐Val‐l ‐Leu‐l ‐Val‐l ‐Ile) ( 1 ) and cyclo(l ‐Leu‐l ‐Leu‐l ‐Ala‐l ‐Ala) ( 2 ), and 15 known compounds, cyclo(Gly‐l ‐Leu‐Gly‐l ‐Leu) ( 3 ), cyclo(l ‐Ser‐l ‐Phe) ( 4 ), cyclo(l ‐Leu‐l ‐Ile) ( 5 ), cyclo(l ‐Tyr‐l ‐Phe) ( 6 ), cyclo(Gly‐l ‐Trp) ( 7 ), cyclo(l ‐Leu‐l ‐Tyr) ( 8 ), cyclo(Gly‐l ‐Phe) ( 9 ), cyclo(l ‐Phe‐trans‐4‐hydroxy‐l ‐Pro) ( 10 ), cyclo(l ‐Leu‐l ‐Leu) ( 11 ), cyclo(l ‐Val‐l ‐Phe) ( 12 ), cyclo(l ‐Val‐l ‐Leu) ( 13 ), cyclo(l ‐Ile‐l ‐Ile) ( 14 ), cyclo(l ‐Tyr‐l ‐Tyr) ( 15 ), turnagainolide A ( 16 ), and bacimethrin ( 17 ) were isolated from the fermentation broth of Streptomyces rutgersensis T009 obtained from Elaphodus davidianus excrement. Their structures were identified on the basis of spectroscopic analysis. Meanwhile, the absolute configurations of the amino acid residues of compounds 1 and 2 were determined by advanced Marfey method. Compound 3 was obtained from a natural source for the first time. The X‐ray single crystal diffraction data of bacimethrin ( 17 ) were also reported for the first time. Compounds 1  –  17 exhibited no antimicrobial activities with the MICs > 100 μg/ml.  相似文献   

18.
The incorporation of the β‐amino acid residues into specific positions in the strands and β‐turn segments of peptide hairpins is being systematically explored. The presence of an additional torsion variable about the C(α) C(β) bond (θ) enhances the conformational repertoire in β‐residues. The conformational analysis of three designed peptide hairpins composed of α/β‐hybrid segments is described: Boc‐Leu‐Val‐Val‐DPro‐β Phe ‐Leu‐Val‐Val‐OMe ( 1 ), Boc‐Leu‐Val‐β Val ‐DPro‐Gly‐β Leu ‐Val‐Val‐OMe ( 2 ), and Boc‐Leu‐Val‐β Phe ‐Val‐DPro‐Gly‐Leu‐β Phe ‐Val‐Val‐OMe ( 3 ). 500‐MHz 1H‐NMR Analysis supports a preponderance of β‐hairpin conformation in solution for all three peptides, with critical cross‐strand NOEs providing evidence for the proposed structures. The crystal structure of peptide 2 reveals a β‐hairpin conformation with two β‐residues occupying facing, non‐H‐bonded positions in antiparallel β‐strands. Notably, βVal(3) adopts a gauche conformation about the C(α) C(β) bond (θ=+65°) without disturbing cross‐strand H‐bonding. The crystal structure of 2 , together with previously published crystal structures of peptides 3 and Boc‐β Phe ‐β Phe ‐DPro‐Gly‐β Phe ‐β Phe ‐OMe, provide an opportunity to visualize the packing of peptide sheets with local ‘polar segments' formed as a consequence of reversal peptide‐bond orientation. The available structural evidence for hairpins suggests that β‐residues can be accommodated into nucleating turn segments and into both the H‐bonding and non‐H‐bonding positions on the strands.  相似文献   

19.
Racemates of hydrophobic amino acids with linear side chains are known to undergo a unique series of solid‐state phase transitions that involve sliding of molecular bilayers upon heating or cooling. Recently, this behaviour was shown to extend also to quasiracemates of two different amino acids with opposite handedness [Görbitz & Karen (2015). J. Phys. Chem. B, 119 , 4975–4984]. Previous investigations are here extended to an l ‐2‐aminobutyric acid–d ‐methionine (1/1) co‐crystal, C4H9NO2·C5H11NO2S. The significant difference in size between the –CH2CH3 and –CH2CH2SCH3 side chains leads to extensive disorder at room temperature, which is essentially resolved after a phase transition at 229 K to an unprecedented triclinic form where all four d ‐methionine molecules in the asymmetric unit have different side‐chain conformations and all three side‐chain rotamers are used for the four partner l ‐2‐aminobutyric acid molecules.  相似文献   

20.
For crystallographic analysis, Leu was substituted for Orn in Gramicidin S (LGS) to suppress interactions with hydrophilic solvent molecules, which increased the flexibility of the Orn side chains, leading to disorder within the crystals. The asymmetric unit (C62H94N10O10·1.296C3H8O·1.403H2O) contains three LGS molecules (A, B and C) forming β‐turn and intramolecular β‐sheet structures. With the exception of one motif in molecule C, d ‐Phe‐Pro turn motifs (Phe is phenylalanine and Pro is proline) were classed as type II′ β‐turns. The peptide backbones twist slightly to the right along the long axis of the molecule. The puckering of Pro is in a Cγ‐endo or twisted Cγ‐endo–Cβ‐exo form. Flanking molecules are arranged such that the angles (A…B = 104°, B…C = 139° and C…A = 117°) form helical β‐sheets. Solvent molecules interact with the peptide backbones supporting the β‐sheets. The forms of the replacement Leu side chains are consistent with the e‐form of the Orn side chain in GS analogues. No hydrophilic region composed of solvent molecules, such as that observed in Gramicidin S hydrochloride (GS·HCl) crystals, was found. The perturbation of αH chemical shifts and coupling constants of CONH showed that the structural properties of GS·HCl and LGS are similar to each other in solution. CD spectra also supported the structural similarity. With the sequence cyclo(–Val–Leu–Leu–d ‐Phe–Pro–)2 (Val is valine and Leu is leucine), LGS lacks the amphiphilicity and antimicrobial activity of parental Gramicidin S (GS). However, the structure of LGS reflects the structural characteristics of GS and no disordering inconvenient for structural analysis was found. Thus, LGS could be a novel scaffold useful for studying β‐turn and sheet structures.  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