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 46 毫秒
1.
Structural prerequisites for the stability of the 31 helix of β-peptides can be defined from inspection of models (Figs. 1 and 2): lateral non-H-substituents in 2- and 3-position on the 3-amino-acid residues of the helix are allowed, axial ones are forbidden. To be able to test this prediction, we synthesized a series of heptapeptide derivatives Boc-(β-HVal-β-HAla-β-HLeu-Xaa-β-HVal-β-HAla-β-HLeu)-OMe 13–22 (Xaa = α- or β-amino-acid residue) and a β-depsipeptide 25 with a central (S)-3-hydroxybutanoic-acid residue (Xaa = –OCH(Me)CH2C(O)–) (Schemes 1 3). Detailed NMR analysis (DQF-COSY, HSQC, HMBC, ROESY, and TOCSY experiments) in methanol solution of the β-hexapeptide H(-β-HVal-β-HAla-β-HLeu)2-OH ( 1 ) and of the β-heptapeptide H-β-HVal-β-HAla-β-HLeu-(S,S)-β-HAla(αMe)-β-HVal-β-HAla- β-HLeu-OH ( 22 ), with a central (2S,3S)-3-amino-2-methylbutanoic-acid residue, confirm the helical structure of such β-peptides (previously discovered in pyridine solution) (Fig.3 and Tables 1–5). The CD spectra of helical β-peptides, the residues of which were prepared by (retentive) Arndt-Eistert homologation of the (S)- or L -α-amino acids, show a trough at 215 nm. Thus, this characteristic pattern of the CD spectra was taken as an indicator for the presence of a helix in methanol solutions of compounds 13–22 and 25 (including partially and fully deprotected forms) (Figs.4–6). The results fully confirm predicted structural effects: incorporation of a single ‘wrong’ residue ((R)-β-HAla, β-HAib, (R,S)-β-HAla(α Me), or N-Me-β-HAla) in the central position of the β-heptapeptide derivatives A (see 17, 18, 20 , or 21 , resp.) causes the CD minimum to disappear. Also, the β-heptadepsipetide 25 (missing H-bond) and the β-heptapeptide analogs with a single α-amino-acid moiety in the middle ( 13 and 14 ) are not helical, according to this analysis. An interesting case is the heptapeptide 15 with the central achiral, unsubstituted 3-aminopropanoic-acid moiety: helical conformation appears to depend upon the presence or absence of terminal protection and upon the solvent (MeOH vs. MeOH/H2O).  相似文献   

2.
Partially and fully protected, and unprotected β-oligopeptides ( 3 – 9 ) were prepared from 1-(aminomethyl)cyclopropanecarboxylic acid, which, in turn, is readily available from cyanoacetate and dibromoethane. N-Boc and C-OMe protection were applied for the fragment-coupling (1-hydroxy-1H-benzotriazole (HOBt)/1-[3-(dimethylamino)propyl]-3-ethylcarbodiimide hydrochloride (EDC)) solution synthesis. X-Ray crystal structures of the dimer ( 3 ), trimer ( 5 ), and tetramer ( 6 ) are described, and compared with those of the Boc-protected building blocks ( 2 ) and of the corresponding trimer ( 10 ) consisting of 1-(aminomethyl)cyclohexanecarbonyl residues (cf. Figs. 1 and 2). While the cyclohexane derivative forms ten-membered hydrogen-bonded rings, the characteristic secondary-structural motif in the cyclopropane derivatives is an eight-membered ring with H-bonding between next neighbors (Fig. 1). All cyclopropanecarbonyl moieties in the reported structures have the – generally more stable – s-cis (`bisected') conformation of the C=O groups on the three-membered rings (not preferred with the cyclohexane analog, the exocyclic CO group of which may be in an s-trans, a perpendicular, an axial, or an equatorial position). The bisecting effect and the large exocyclic bond angle (120°) in the cyclopropane units are proposed to provide the `ordering' elements – on top of the staggering effect of the C(2)−C(3) ethane bond in all β-peptides – which lead to the observed substituent-induced turn formation. A high degree of intramolecular H-bonding is evident also from IR spectroscopy (Fig. 3), and concentration- and temperature-dependent NMR measurements (Fig. 4) of CHCl3 and CD2Cl2 solutions, indicating that the boat-type arrangement of the eight-membered rings with their unusual H-bonding geometry (Fig. 1, f ) is also present in solution. A possible structure of a poly[1-(aminomethyl)cyclopropane-carboxylic acid] consisting of a flight of stairs formed by folded H-bonded eight-membered rings is modelled, using the oligomer X-ray data (Fig. 5). The type of secondary structure found in the solid state of the β2,2-peptides reported here is unprecedented in the realm of α-peptides and proteins.  相似文献   

3.
all-L -β3-Penta-, hexa-, and heptapeptides with the proteinogenic side chains of valine, leucine, serine, cysteine, and methionine have been prepared by previously described procedures ( 12 , 13 , 14 , 15 ; Schemes 2 – 5). Thioether cleavage with Na/NH3 in β-HMet residues has also provided a β3-hexapeptide with homocysteine (CH2CH2S) side chains ( 13e ). The HS−(CH2)n groups were positioned on the β-peptidic backbone in such a way that, upon disulfide-bridge formation, the corresponding β-peptide was expected to maintain either a 31-helical secondary structure ( 1 , 2 ) (Fig. 1) or to be forced to adopt another conformation ( 3 , 4 ). The 13-, 17-, 19-, and 21-membered-ring macrocyclic disulfide derivatives and their open-chain precursors, as well as all synthetic intermediates, were purified (crystallization, flash or preparative HPL chromatography; Fig. 5) and fully characterized (m.p., [α]D, CD, IR, NMR, FAB or ESI mass spectroscopy, and elemental analysis, whenever possible; Fig. 2 and Exper. Part). The structures in MeOH and H2O of the new β-peptides were studied by CD spectroscopy (Figs. 3 and 4), where the characteristic 215-nm-trough/200-nm-peak pattern was used as an indicator for the presence or absence of (M)-31-helical conformations. A CH2−S2−CH2 and, somewhat less so, a (CH2)2−S2−(CH2)2 bracket between residues i and i+3 ( 1 vs. 12d , and 2 vs. 13e in Fig. 3) give rise to CD spectra which are compatible with the presence of 31-helical structures, while CH2−S2−CH2 brackets between residues i and i+2 ( 3 vs. 14c ) or i and i+4 ( 4 vs. 15c in Fig. 4) do not.  相似文献   

4.
The recently improved conditions for solid‐phase synthesis of β3‐peptides by the Fmoc strategy were used to synthesize a β‐tetracosapeptide ( 4 , Scheme) composed of eight different β‐amino acid residues; 11 of the 24 residues carry functionalized proteinogenic side chains (namely those of Glu, Lys, Ser, and Tyr). The highly H2O‐soluble β‐tetracosapeptide was identified by 1H‐NMR spectroscopy (in MeOH), analytical HPL chromatography, and ESI‐mass spectrometry (Fig. 1). The expected 314‐helical secondary structure of the new β‐peptide was designed to have one hydrophobic and two hydrophilic faces, and to be compared with other β‐peptides ( 1 – 3 ), two of which are also of amphipathic character in this secondary structure (Fig. 2). In the absence of NMR‐structural proof, the CD spectra of the four β‐peptides were compared (Figs. 3 and 4). The β‐tetracosapeptide exhibits an unprecedented CD pattern (in MeOH and in H2O solution) that may arise from a new type of secondary structure or from an unordered conformation.  相似文献   

5.
β‐Hexa‐, β‐hepta‐, and β‐nonapeptides, 1 – 6 , which carry functionalized side chains (CO2R, CO, (CH2)4NH, CH2−CH=CH2) consisting of β3‐amino‐acid residues of alternating configuration, or which carry geminal substituents in the 2‐ or 3‐positions of all residues, have been synthesized (Schemes 1 – 3), and their CD spectra in MeOH are reported (Figs. 2 – 6). Strong Cotton effects (Θ>105) are indicative of the presence of chiral secondary structures. It is suggested by simple modelling (Fig. 1) that the new β‐peptides should not be able to fold to the familiar 314‐helical structures. Still, three of them ( 3 , 4 , and 5 ) give rise to CD spectra matching those of β‐peptides that are known to be present as (M)‐ or (P)‐314‐helices in MeOH solution. While possible folding motifs (Figs. 3,b, and 7) of the new β‐peptides have been identified in crystal structures, an interpretation of the CD spectra has to be postponed until NMR solution structures become available. A list of all β‐peptides giving rise to CD spectra with a minimum near 215 nm is included (Table).  相似文献   

6.
An icosapeptide, 1 , containing the β3‐amino acid residues with the 20 proteinogenic side chains has been assembled by manual solid‐phase synthesis, according to the Fmoc strategy. The sequence was chosen in such a way that a possible 314‐helical conformation (secondary structure) would be stabilized by salt bridges and have an amphipathic character (Fig. 1,a), and the N‐terminal β3hCys would lend itself to thioligations and disulfide formation ( 2 and 3 , in Figs. 1 and 2). The products 1 – 3 were pure according to RP‐HPLC, NMR, and MS analysis (Fig. 1,b and c, Fig. 2,c and d, and Fig. 3). With due caution, the CD spectra in aqueous solution (pH 7) and in MeOH (Fig. 4), with normalized Cotton effects θ =?14000 to ?16000 [deg?cm2?dmol?1] between 209 and 210 nm, might be taken as an evidence for the presence of 314‐helical conformations. An evaluation of the data from a 700‐MHz 2D‐NMR measurement of the disulfide 2 in CD3OH is in progress.  相似文献   

7.
The stability constants of acetylcholinium, carbamoylcholinium, and cholinium diiodochlorides and diiodobromides in chloroform solutions were determined and the kinetics of disproportionation of these systems in 1 : 9 CHCl3—MeOH (MeCN) mixtures were studied by UV spectroscopy. A possible mechanism of mutual transformations of the polyhalides is proposed and an interrelation between the nature of the iodine-coordinating solvent and the extent of reversibility of the process is established. The electronic structures and relative stabilities of acetylcholinium iodohalides and charge-transfer complexes S·XI2 and S·I2 (S = MeOH, MeCN, CHCl3; X = Cl, Br, I) were studied by ab initio RHF and MP2(full) methods in the HW+(3d) and 6-31G++(d,p) basis sets. It was found that all the solvents studied favor the decomposition of the iodohalide anions to liberate molecular iodine; however, disproportionation of I2 is possible only for the S·I2 complexes with a high extent of charge transfer.  相似文献   

8.
The crystal structure of (±)-7,8,13β,14α-tetrahydro-N7-(13C)methylcorysaminium iodide (13C- 3a ·I) was investigated by X-ray analysis and thus the relative configuration (7S*,13S*,14S*) established (Fig. 1). The conformation of 3a was shown to have a cis-junction of the B/C rings and the rings A and D in an antiperiplanar position relative to the C(13)? C(14) bond (‘anti-cis’), a twisted half-chair for ring B, and a half-chair for ring C (Figs. 2 and 3). Conformation analysis by 1H-NMR data indicated that the crystal conformation of 3a is also the preferred one in MeOH solution.  相似文献   

9.
β‐Peptides offer the unique possibility to incorporate additional heteroatoms into the peptidic backbone (Figs. 1 and 2). We report here the synthesis and spectroscopic investigations of β2‐peptide analogs consisting of (S)‐3‐aza‐β‐amino acids carrying the side chains of Val, Ala, and Leu. The hydrazino carboxylic acids were prepared by a known method: Boc amidation of the corresponding N‐benzyl‐L ‐α‐amino acids with an oxaziridine (Scheme 1). Couplings and fragment coupling of the 3‐benzylaza‐β2‐amino acids and a corresponding tripeptide (N‐Boc/C‐OMe strategy) with common peptide‐coupling reagents in solution led to β2‐di, β2‐tri‐, and β2‐hexaazapeptide derivatives, which could be N‐debenzylated ( 4 – 9 ; Schemes 2–4). The new compounds were identified by optical rotation, and IR, 1H‐ and 13C‐NMR, and CD spectroscopy (Figs. 4 and 5) and high‐resolution mass spectrometry, and, in one case, by X‐ray crystallography (Fig. 3). In spite of extensive measurements under various conditions (temperatures, solvents), it was not possible to determine the secondary structure of the β2‐azapeptides by NMR spectroscopy (overlapping and broad signals, fast exchange between the two types of NH protons!). The CD spectra of the N‐Boc and C‐OMe terminally protected hexapeptide analog 9 in MeOH and in H2O (at different pH) might arise from a (P)‐314‐helical structure. The N‐Boc‐β2‐tri and N‐Boc‐β2‐hexaazapeptide esters, 7 and 9 , were shown to be stable for 48 h against the following peptidases: pronase, proteinase K, chymotrypsin, trypsin, carboxypeptidase A, and 20S proteasome.  相似文献   

10.
An all‐β3‐dodecapeptide with a protected N‐terminal thiol‐anchoring group and with seven side chains has been synthesized in multi‐mg amounts by the manual solid‐phase technique, applying Fmoc methodology and the Wang resin. The sequence is β‐HLys‐β‐HPhe‐β‐HTyr‐β‐HLeu‐β‐HLys‐β‐HSer‐β‐HLys‐β‐HPhe‐β‐HSer‐β‐HVal‐β‐HLys‐β‐HAla‐OH (from N‐ to C‐terminus; see 1 ). The functional groups in the side chains of the building blocks were Boc (β‐HLys) or t‐Bu ether (β‐HSer, β‐HTyr) protected to allow for simultaneous deprotection and detachment from the resin with trifluoroacetic acid. All coupling steps were achieved with HBTU (=O‐(1H‐benzotriazol‐1‐yl)‐1,1,3,3‐tetramethyl uronium hexafluorophosphate)/HOBt (=1‐hydroxy‐1H‐benzotriazole) in DMF. For Fmoc (=(9H‐fluoren‐9‐yl)methoxycarbonyl) deprotection, a protocol was developed to surmount the previously reported problems arising in solid‐phase synthesis of β‐peptides when the chain length exceeds seven or eight amino‐acid moieties: for up to seven amino acids, a 20% solution of piperidine in DMF was used for removal of Fmoc; for the subsequent five amino acids, DBU and piperidine were employed for complete deprotection. The crude product was purified by preparative reversed‐phase HPLC, and the yield of pure β‐dodecapeptide derivative ( 1 ) was 23%. As the compound is well‐soluble in H2O, it was characterized by 1H‐NMR (in MeOH and H2O), 13C‐NMR (in MeOH), and CD spectroscopy (in MeOH and in H2O at pH values ranging from 3.5 to 11), and its molecular weight and composition were confirmed by high‐resolution mass spectrometry (Figs. 1 – 4). In MeOH solution, the β‐dodecapeptide exhibits the expected CD pattern typical of an (M)‐314‐helical secondary structure. In H2O, however, the characteristic trough near 215 nm is missing in the CD spectrum, only a strong positive Cotton effect at 202 nm was observed, indicating the presence of β‐peptidic secondary structures, containing ten‐membered H‐bonded rings, such as the 12/10 helix (Fig. 4, right) or the hairpin. Only a detailed NMR solution‐structure analysis will provide the clues necessary for understanding the effects leading to the observed dramatic structural change of the highly functionalized β‐dodecapeptide described.  相似文献   

11.
The GROMOS96 molecular‐dynamics (MD) program and force field was used to calculate the conformations at 298 K in CHCl3 solution of two hexakis(3‐hydroxyalkanoic acids). One consists of (R)‐3‐hydroxybutanoate (HB) residues only: H−(OCH(Me)−CH2−CO)6−OH ( 1 ). The other one carries the side chains of valine, alanine, and leucine: H−(OCH(CHMe2)CH2−CO−O−CH(Me)−CH2−CO−O−CH(CH2 CHMe2)−CH2−CO)2−OH ( 2 ), with homochiral 3‐hydroxyalkanoate (HA) moieties. In both cases, the conformational equilibria were sampled 2500 times for 25 ns. Other than clusters of arrangements with inter‐residual hydrogen bonding (between the O‐ and C‐terminal OH and COOH groups, and with chain‐bound ester carbonyl O‐atoms; Fig. 6), there are no preferred backbone conformations in which the molecules 1 and 2 spend more than 5% of the time. Specifically, neither the 21‐ nor the 31‐helical conformation of the oligoester backbone (found in stretched fibers, in lamellar crystallites, and in single crystals of polymers PHB and of oligomers OHB) is sampled to any significant extent (Fig. 8 and 9), in spite of the high population, in both oligomers, of the (−)‐synclinal conformation around the C(2)−C(3) bond (angle ϕ2 in Fig. 2). In contrast to β‐oligopeptides, for which strongly preferred secondary structures are found after a few ns, and for which the number of conformations levels off with time, the number of conformational clusters of the corresponding oligoesters found by our force‐field MD calculations increases steadily over the observation time of 25 ns (Fig. 5). Thus, the conclusion from biological and physical‐chemical studies, according to which the PHB chain is extremely flexible, is confirmed by our computational investigation: in CHCl3 solution, the hexakis(3‐hydroxyalkanoate) chain samples its conformational space randomly!  相似文献   

12.
The correlation between β2‐, β3‐, and β2,3‐amino acid‐residue configuration and stability of helix and hairpin‐turn secondary structures of peptides consisting of homologated proteinogenic amino acids is analyzed (Figs. 1–3). To test the power of Zn2+ ions in fortifying and/or enforcing secondary structures of β‐peptides, a β‐decapeptide, 1 , four β‐octapeptides, 2 – 5 , and a β‐hexadecapeptide, 10 , have been devised and synthesized. The design was such that the peptides would a) fold to a 14‐helix ( 1 and 3 ) or a hairpin turn ( 2 and 4 ), or form neither of these two secondary structures (i.e., 5 ), and b) carry the side chains of cysteine and histidine in positions, which will allow Zn2+ ions to use their extraordinary affinity for RS? and the imidazole N‐atoms for stabilizing or destabilizing the intrinsic secondary structures of the peptides. The β‐hexadecapeptide 10 was designed to a) fold to a turn, to which a 14‐helical structure is attached through a β‐dipeptide spacer, and b) contain two cysteine and two histidine side chains for Zn complexation, in order to possibly mimic a Zn‐finger motif. While CD spectra (Figs. 6–8 and 17) and ESI mass spectra (Figs. 9 and 18) are compatible with the expected effects of Zn2+ ions in all cases, it was shown by detailed NMR analyses of three of the peptides, i.e., 2, 3, 5 , in the absence and presence of ZnCl2, that i) β‐peptide 2 forms a hairpin turn in H2O, even without Zn complexation to the terminal β3hHis and β3hCys side chains (Fig. 11), ii) β‐peptide 3 , which is present as a 14‐helix in MeOH, is forced to a hairpin‐turn structure by Zn complexation in H2O (Fig. 12), and iii) β‐peptide 5 is poorly ordered in CD3OH (Fig. 13) and in H2O (Fig. 14), with far‐remote β3hCys and β3hHis residues, and has a distorted turn structure in the presence of Zn2+ ions in H2O, with proximate terminal Cys and His side chains (Fig. 15).  相似文献   

13.
The preparation of (S)‐β2,2,3‐amino acids with two Me groups in the α‐position and the side chains of Ala, Val, and Leu in the β‐position (double methylation of Boc‐β‐HAla‐OMe, Boc‐β‐Val‐OMe, and Boc‐β‐Leu‐OMe, Scheme 2) is described. These β‐amino acids and unlabelled as well as specifically 13C‐ and 15N‐labelled 2,2‐dimethyl‐3‐amino acid (β2,2‐HAib) derivatives have been coupled in solution (Schemes 1, 3 and 4) to give protected (N‐Boc, C‐OMe), partially protected (N‐Boc/C‐OH, N‐H/C‐OMe), and unprotected β2,2‐ and β2,2,3‐hexapeptides, and β2,2‐ and β2,2,3‐heptapeptides 1 – 7 . NMR Analyses in solution (Tables 1 and 2, and Figs. 2–4) and in the solid state (2D‐MAS NMR measurements of the fully labelled Boc‐(β2,2‐HAib)6‐OMe ([13C30, 15N6]‐ 1e ; Fig. 5), and TEDOR/REDOR NMR investigations of mixtures (Fig. 6) of the unlabelled Ac‐(β2,2‐HAib)7‐OMe ( 4 ) and of a labelled derivative ([13C4,15N2]‐ 5 ; Figs. 7–11, and 19), a molecular‐modeling study (Figs. 13–15), and a search in the Cambridge Crystallographic Data Base (Fig. 16) allow the following conclusions: i) there is no evidence for folding (helix or turn) or for aggregation to sheets of the geminally dimethyl substituted peptide chains in solution; ii) there are distinct conformational preferences of the individual β2,2‐ and β2,2,3‐amino acid residues: close to eclipsing around the C(O) C(Me2(CHR)) bond (τ1,2), almost perfect staggering around the C(2) C(3) ethane bond (τ2,3), and antiperiplanar arrangement of H(C3) and H(N) (τ3,N; Fig. 12) in the solid state; iii) the β2,2‐peptides may be part of a turn structure with a ten‐membered H‐bonded ring; iv) the main structure present in the solid state of F3CCO(β2,2‐HAib)7‐OMe is a nonfolded chain (>30 Å between the termini and >20 Å between the N‐terminus and the CH2 group of residue 5) with all CO bonds in a parallel alignment (±10°). With these structural parameters, a simple modelling was performed producing three (maybe four) possible chain geometries: one fully extended, two with parallel peptide planes (with zick‐zack and crankshaft‐type arrangement of the peptide bonds), and (possibly) a fourth with meander‐like winding ( D – G in Figs. 17 and 18).  相似文献   

14.
The β-hexapeptide (H-β-HVal-β-HAla-β-HLeu)2-OH ( 2 ) was prepared from the component L -β-amino acids by conventional peptide synthesis, including fragment coupling. A cyclo-β-tri- and a cyclo-β-hexapeptide were also prepared. The β-amino acids were obtained from α-amino acids by Arndt-Eistert homologation. All reactions leading to the β-peptides occur smoothly and in high yields. The β-peptides were characterized by their CD and NMR spectra (COSY, ROESY, TOCSY, and NOE-restricted modelling), and by an X-ray crystal-structure analysis. β-Sheet-type structures (in the solid state) and a compact, left-handed or (M) 31 helix of 5-Å pitch (in solution) were discovered. Comparison with the analogous secondary structures of α-peptides shows fundamental differences, the most surprising one at this point being the greater stability of β-peptide helices. There are structural relationships of β-peptides with oligomers of β-hydroxyalkanoic acids, and dissimilarities between the two classes of compounds are a demonstration of the power of H-bonding. The β-hexapeptide 2 is stable to cleavage by pepsin at pH 2 in H2O for at least 60 h at 37°, while the corresponding α-peptide H-(Val-Ala-Leu)2-OH is cleaved instantaneously under these conditions. The implication of the described results are discussed.  相似文献   

15.
通过微波"一锅法"合成了4个双核苄基锡配合物:{[C_4H_3S(O)C=N-N=C(Me)COO](PhCH_2)_2Sn(MeOH)}_2(C1)、{[C_4H_3S(O)C=NN=C (Me)COO](p-Cl-C_6H_4CH_2)_2Sn (MeOH)}_2(C2)、{[C_4H_3S (O)C=N-N=C (PhCH_2)COO](PhCH_2)_2Sn (MeOH)}_2(C3)、{[C_4H_3S (O)C=N-N=C(PhCH_2)COO](p-Cl-C_6H_4CH_2)_2Sn(MeOH)}_2(C4),利用元素分析、IR、~1H NMR、~(13)C NMR、~(119)Sn NMR、HRMS以及X射线单晶衍射等表征了配合物结构。4个配合物分子均为双锡核分子,以Sn_2O_2四元环为中心对称,且中心锡原子与配位原子形成七配位畸变五角双锥构型。测试了配合物C1~C4的热稳定性以及配合物对癌细胞H460、HepG2、MCF7的体外抑制活性,结果表明:配合物C2是4个新合成的配合物中抑制癌细胞效果最好的化合物。  相似文献   

16.
Selective lipid transport through the brush-border membrane in the small intestine of mammals is mediated by membrane-bound proteins, the so-called scavenger receptors of class B, type I or II (SR-BI or -BII). These, in turn, are inhibited by certain proteins and synthetic α-peptides that have an amphipathic helix as the binding motif (Fig. 1). In whole cells (test with human colonic carcinoma cells, CaCo-2), on the other hand, the inhibitors are subject to proteolysis. We have now tested six β-peptides (hexa-, hepta-, and nonamers 1 – 6 ), each carrying one to seven water-solubilizing side chains of either Ser or Lys, with a brush-border-membrane (BBM) vesicle model system (rate and IC50 values in Figs. 2 and 3) and with a tightly packed monolayer of CaCo-2 cells (rate in Fig. 4), to find that the rate of transport of cholesterol can be reduced to what may be considered the passive diffusion (`background') level. There is a correlation between the ability of the β-peptides to form an amphipathic-type 314-helical secondary structure in MeOH and their inhibitory effect (Table 1 and Fig. 5). Although the inhibitory activity of the β-peptides is in only the mM range (Table 2), it is to be compared with no activity at all of previously tested α-peptides and proteins (built of L -amino acids) in CaCo-2 cells. Furthermore, these active β-peptides ( 1 , 5 , and 6 ) contain only seven or nine residues and must be considered simple, first-generation models capable of mimicking the biological activity of amphipathic α-peptide helices in living whole cells.  相似文献   

17.
β-Hexapeptides 1–5 and a β-dodecapeptide 6 with sequences containing two different types of β-amino acids (aliphatic proteinageous side chains in the 2- or in the 3-position) have been prepared. CD (Fig. 1) and NMR measurements indicate that, with one exception, the secondary structures formed by these new β-peptides differ from those of isomers studied previously. Detailed NMR analysis of the β-hexapeptide 5 (with alternating β23-building blocks) and molecular-dynamics simulations have produced a minimum energy conformation (Fig. 2,b)which might be described as a novel irregular helix containing ten- and twelve-membered H-bonded rings. This demonstrates the great structural variability of β-peptides, since three different helical secondary structures have been discovered to date.  相似文献   

18.
We report on the synthesis of new and previously described β-peptides ( 1 – 6 ), consisting of up to twelve β2,2- or β3,3-geminally disubstituted β-amino acids which do not fit into any of the secondary structural patterns of β-peptides, hitherto disclosed. The required 2,2- and 3,3-dimethyl derivatives of 3-aminopropanoic acid are readily obtained from 3-methylbut-2-enoic acid and ammonia (Scheme 1) and from Boc-protected methyl 3-aminopropanoate by enolate methylation (Scheme 2). Protected (Boc for solution-, Fmoc for solid-phase syntheses) 1-(aminomethyl)cycloalkanecarboxylic-acid derivatives (with cyclopropane, cyclobutane, cyclopentane, and cyclohexane rings) are obtained from 1-cyanocycloalkanecarboxylates and the corresponding dihaloalkanes (Scheme 3). Fully 13C- and 15N-labeled 3-amino-2,2-dimethylpropanoic-acid derivatives were prepared from the corresponding labeled precursors (see asterixed formula numbers and Scheme 4). Coupling of these amino acids was achieved by methods which we had previously employed for other β-peptide syntheses (intermediates 18 – 23 ). Crystal structures of Boc-protected geminally disubstituted amino acids ( 16a – d ) and of the corresponding tripeptide ( 23a ), as well as NMR and IR spectra of an isotopically labeled β-hexapeptide ( 2a* ) are presented (Figs. 1 – 4) and discussed. The tripeptide structure contains a ten-membered H-bonded ring which is proposed to be a turn-forming motif for β-peptides (Fig. 2).  相似文献   

19.
The triolide of (R)-3-hydroxybutanoic acid ((R,R,R,))-3,7,11-trimethyl-2,6,10-trioxadodecane-1,5,9-trione; ( 1 ), readily available from the corresponding biopolymer P(3-HB) in one step, forms crystalline complexes with alkali and alkaline earth salts. The X-ray crystal structures of three such complexes, (3 NaSCN)·4 1 ( 2 ), (2 KSCN)·2 1 · H2O ( 3 ), and (2) Ba(SCN)2 · 2 1 · 2 H2O · THF ( 4 ), have been determined and are compared. The triolide is found in these structures (i) as a free molecule, making no contacts with a cation (clathrate-type inclusion), (ii) as a monodentate ligand coordinated to a single ion with one carbonyl O-atom only, (iii) as a chelator, forming an eight-membered ring, with two carbonyl O-atoms attached to the same ion, (iv) as a linker, using two carbonyl O-atoms to bind to the two metals of an ion-X-ion unit (ten-membered ring), and (v), in a crown-ester complex, in which an ion is sitting on the three unidirectional C?O groups of a triolide molecule (Figs. 1–3). The crystal packing is such that there are columns along certain axes in the centers of which the cations are surrounded by counterions and triolide molecules, with the non-polar parts of 1 on the outside (Fig. 4). In the complexes 2–4 , the triolide assumes conformations which are slightly distorted, with the carbonyl O-atoms moved closer together, as compared to the ‘free’ triolide 1 (Fig. 5). These observed features are compatible with the view that oligo (3-HB) may be involved in the formation of Ca polyphosphate ion channels through cell membranes. A comparison is also made between the triolide structure in 1–4 and in enterobactin, a super Fe chelator (Fig. 5). To better understand the binding between the Na ion and the triolide carbonyl O-atoms in the crown-ester complex, we have applied electron-localization function (ELF) calculations with the data set of structure 2 , and we have produced ELF representations of ethane, ethene, and methyl acetate (Figs. 6–9). It turns out that this theoretical method leads to electron-localization patterns which are in astounding agreement with qualitative bonding models of organic chemists, such as the ‘double bond character of the CO? OR single bond’ or the ‘hyperconjugative n → σ* interactions between lone pairs on the O-atoms and neighbouring σ-bonds’ in ester groups (Fig. 8). The noncovalent, dipole/pole-type character of bonding between Na+ and the triolide carbonyl O-atoms in the crown-ester complex (the Na? O?C plane is roughly perpendicular to the O? C?O plane) is confirmed by the ELF calculation; other bonding features such as the C?N bond in the NaSCN complex 2 are also included in the discussion (Fig. 9).  相似文献   

20.
Aervins A‐D (1‐4), four new coumaronochromone analogues have been isolated from the CHCl3‐soluble fraction of the MeOH extract of the whole plant of Aerva persica. Their structures were assigned based on 1H NMR, 13C NMR spectra, DEPT, and by 2DNMR experiments. Copyright © 2009 John Wiley & Sons, Ltd.  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