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.
γ4‐Tripeptides and γ4‐hexapeptides, 1 – 4 , with OH groups in the 2‐ or 3‐position on each residue have been prepared. The corresponding 2‐hydroxy amino acids were obtained by Si‐nitronate (3+2) cycloadditions to the acryloyl derivative of Oppolzer's sultam and Raney‐Ni reduction of the resulting 1,2‐oxazolidines (Scheme 1). The 3‐hydroxy amino acid derivatives were prepared by chain elongation via Claisen condensation of Boc‐Ala‐OH, Boc‐Val‐OH, and Boc‐Leu‐OH, and NaBH4 reduction of the methyl 4‐amino 3‐oxo carboxylates formed (Scheme 2). The N‐Boc hydroxy amino acids were coupled in solution to give the γ‐peptides. CD Spectra of the new types of γ‐peptides were recorded and compared with those of simple γ2‐, γ3‐, γ4‐, and γ2,3,4‐peptides (Figs. 3, 4, and 5). An intense Cotton effect at ca. 200 nm ([Θ]=−2⋅105 deg⋅cm2⋅dmol−1) indicates that the hexapeptide built of (3R,4S)‐4‐amino‐3‐hydroxy acids (with the side chains of Val, Ala, Leu) folds to a secondary structure so far unknown. The stability of peptides from β‐ and γ‐amino acids, which carry heteroatoms on their backbones is discussed (Fig. 1). Positions on the γ‐peptidic 2.614 helix are identified at which non‐H‐atoms are `allowed' (Fig. 2).  相似文献   

2.
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   

3.
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   

4.
Hybrid peptides composed of α‐ and β‐amino acids have recently emerged as new class of peptide foldamers. Comparatively, γ‐ and hybrid γ‐peptides composed of γ4‐amino acids are less studied than their β‐counterparts. However, recent investigations reveal that γ4‐amino acids have a higher propensity to fold into ordered helical structures. As amino acid side‐chain functional groups play a crucial role in the biological context, the objective of this study was to investigate efficient synthesis of γ4‐residues with functional proteinogenic side‐chains and their structural analysis in hybrid‐peptide sequences. Here, the efficient and enantiopure synthesis of various N‐ and C‐terminal free‐γ4‐residues, starting from the benzyl esters (COOBzl) of N‐Cbz‐protected (E)α,β‐unsaturated γ‐amino acids through multiple hydrogenolysis and double‐bond reduction in a single‐pot catalytic hydrogenation is reported. The crystal conformations of eight unprotected γ4‐amino acids (γ4‐Val, γ4‐Leu, γ4‐Ile, γ4‐Thr(OtBu), γ4‐Tyr, γ4‐Asp(OtBu), γ4‐Glu(OtBu), and γ‐Aib) reveals that these amino acids adopted a helix favoring gauche conformations along the central Cγ? Cβ bond. To study the behavior of γ4‐residues with functional side chains in peptide sequences, two short hybrid γ‐peptides P1 (Ac‐Aib‐γ4‐Asn‐Aib‐γ4‐Leu‐Aib‐γ4‐Leu‐CONH2) and P2 (Ac‐Aib‐γ4‐Ser‐Aib‐γ4‐Val‐Aib‐γ4‐Val‐CONH2) were designed, synthesized on solid phase, and their 12‐helical conformation in single crystals were studied. Remarkably, the γ4‐Asn residue in P1 facilitates the tetrameric helical aggregations through interhelical H bonding between the side‐chain amide groups. Furthermore, the hydroxyl side‐chain of γ4‐Ser in P2 is involved in the interhelical H bonding with the backbone amide group. In addition, the analysis of 87 γ4‐residues in peptide single‐crystals reveal that the γ4‐residues in 12‐helices are more ordered as compared with the 10/12‐ and 12/14‐helices.  相似文献   

5.
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   

6.
While the chain conformation of poly‐ and oligo[(R)‐3‐hydroxybutanoate] (PHB, OHB) is known to be 21‐ and 31‐helical in stretched fibers and in the crystalline state, respectively (Fig. 2), the structure in solution is unknown. To be able to determine the NMR‐solution structure, specifically labeled linear oligomers have been prepared: a 16‐mer consisting of alternating pairs of fully 13C‐labeled and non‐labeled residues ( 1 ) and a 20‐mer containing an O13CH(13CH2D)‐13CHDSi13CO residue in position 9 (from the O‐terminus) and a fully 13C‐labeled residue in position 12 ( 2 ), both with (t‐Bu)Ph2Si protection at the O‐ and Bn protection at the C‐terminus. The labeled (R)‐3‐hydroxybutanoic acid building blocks were prepared by Noyori hydrogenation of the ethyl ester of fully 13C‐labeled acetoacetic acid, and the D‐atoms were incorporated by D2/Pd‐C reduction of a previously reported dibromo‐1,3‐dioxinone 8 (Scheme 1). The oligomers were obtained by a series of fragment couplings (Schemes 2 and 3). 600‐MHz NMR COSY, HSQC, ROESY, and cross‐correlated relaxation measurements (Figs. 46, 9, and 12, and Tables 13) at different temperatures and interpretations thereof led to assignments of all resonances, including those from the diastereotopic C(2)H2 protons, and to determination of the conformationally averaged dihedral angles ϕ2 and ϕ3 (Figs. 2, 7, and 8) in the chain of the oligoester. The conclusions are: all but five or six terminal residues adopt the same conformation; the 21 helix is not the predominant secondary structure; the structure of the HB chain is averaged, even at –30°. Our investigation confirms the high flexibility of the polyester chain, a property that has been deduced previously from biological studies of PHB in membranes, in ion channels, and as appendage of proteins.  相似文献   

7.
α,β‐Unsaturated amino acids (dehydroamino acids) have been found in naturally occurring antibiotics of microbial origin and in some proteins. Due to the presence of the CαCβ double bond, the dehydroamino acids influence the main‐chain and the side‐chain conformations. The lowest‐energy conformational state of the model tripeptides, Ac–X–ΔAla–NHMe, (X=Ala, Val, Leu, Abu, or Phe) corresponds to ϕ1=−30°, ψ1=120° and ϕ22=30°. This structure is stabilized by the hydrogen bond between CO of the acetyl group and the NH of the amide group, resulting in the formation of a 10‐membered ring. In the model heptapeptide containing ΔAla at alternate position with Ala, Abu, and Leu, the lowest‐energy conformation corresponds to ϕ=−30° and ψ=120° for all the Ala, Abu, and Leu residues and ϕ=ψ=30° for all ΔAla residues. A graphical view of the molecule in this conformation reveals the formation of three hydrogen bonds involving the CO moiety of the ith residue and the NH moiety of the i+3th residue, resulting in a 10‐membered ring formation. In this structure, only alternate peptide bonds are involved in the intramolecular hydrogen‐bond formation unlike the helices and it has been named the β‐bend ribbon structure. The helical structures were predicted to be the most stable structures in the heptapeptide Ac–(Aib–ΔAla)3–NHMe with ϕ=±30°, ψ=±60° for Aib residues and ϕ=ψ=±30° for ΔAla residues. The computational results reveal that the ΔAla residue does not induce an inverse γ‐turn in the preceding residue. It is the competitive interaction of small solvent molecules with the hydrogen‐bonding sites of the peptide which gives rise to the formation of an inverse γ‐turn (ϕ1=−54°, ψ1=82°; ϕ2=44°, ψ2=3°) in the preceding residue to ΔAla. The computational studies for the positional preference of ΔAla in the peptide containing one ΔAla and nine Ala residues reveals the formation of a 310 helical structure in all the cases with the terminal preferences for ΔAla, consistent with the position of ΔAla in the natural antibiotics. The extended structures is found to be the most stable for poly‐ΔAla. ©1999 John Wiley & Sons, Inc. Int J Quant Chem 72: 15–23, 1999  相似文献   

8.
The preparation of (2S,3S)‐ and (2R,3S)‐2‐fluoro and of (3S)‐2,2‐difluoro‐3‐amino carboxylic acid derivatives, 1 – 3 , from alanine, valine, leucine, threonine, and β3h‐alanine (Schemes 1 and 2, Table) is described. The stereochemical course of (diethylamino)sulfur trifluoride (DAST) reactions with N,N‐dibenzyl‐2‐amino‐3‐hydroxy and 3‐amino‐2‐hydroxy carboxylic acid esters is discussed (Fig. 1). The fluoro‐β‐amino acid residues have been incorporated into pyrimidinones ( 11 – 13 ; Fig. 2) and into cyclic β‐tri‐ and β‐tetrapeptides 17 – 19 and 21 – 23 (Scheme 3) with rigid skeletons, so that reliable structural data (bond lengths, bond angles, and Karplus parameters) can be obtained. β‐Hexapeptides Boc[(2S)‐β3hXaa(αF)]6OBn and Boc[β3hXaa(α,αF2)]6‐OBn, 24 – 26 , with the side chains of Ala, Val, and Leu, have been synthesized (Scheme 4), and their CD spectra (Fig. 3) are discussed. Most compounds and many intermediates are fully characterized by IR‐ and 1H‐, 13C‐ and 19F‐NMR spectroscopy, by MS spectrometry, and by elemental analyses, [α]D and melting‐point values.  相似文献   

9.
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   

10.
The preparation of three new N‐Fmoc‐protected (Fmoc=[(9H‐fluoren‐9‐yl)methoxy]carbonyl) β2‐homoamino acids with proteinogenic side chains (from Ile, Tyr, and Met) is described, the key step being a diastereoselective amidomethylation of the corresponding Ti‐enolates of 3‐acyl‐4‐isopropyl‐5,5‐diphenyloxazolidin‐2‐ones with CbzNHCH2OMe/TiCl4 (Cbz=(benzyloxy)carbonyl) in yields of 60–70% and with diastereoselectivities of >90%. Removal of the chiral auxiliary with LiOH or NaOH gives the N‐Cbz‐protected β‐amino acids, which were subjected to an N‐Cbz/N‐Fmoc (Fmoc=[(9H‐fluoren‐9‐yl)methoxy]carbonyl) protective‐group exchange. The method is suitable for large‐scale preparation of Fmoc‐β2hXaa‐OH for solid‐phase syntheses of β‐peptides. The Fmoc‐amino acids and all compounds leading to them have been fully characterized by melting points, optical rotations, IR, 1H‐ and 13C‐NMR, and mass spectra, as well as by elemental analyses.  相似文献   

11.
Two new series of Boc‐N‐α,δ‐/δ,α‐ and β,δ‐/δ,β‐hybrid peptides containing repeats of L ‐Ala‐δ5‐Caa/δ5‐Caa‐L ‐Ala and β3‐Caa‐δ5‐Caa/δ5‐Caa‐β3‐Caa (L ‐Ala = L ‐alanine, Caa = C‐linked carbo amino acid derived from D ‐xylose) have been differentiated by both positive and negative ion electrospray ionization (ESI) ion trap tandem mass spectrometry (MS/MS). MSn spectra of protonated isomeric peptides produce characteristic fragmentation involving the peptide backbone, the Boc‐group, and the side chain. The dipeptide positional isomers are differentiated by the collision‐induced dissociation (CID) of the protonated peptides. The loss of 2‐methylprop‐1‐ene is more pronounced for Boc‐NH‐L ‐Ala‐δ‐Caa‐OCH3 (1), whereas it is totally absent for its positional isomer Boc‐NH‐δ‐Caa‐L ‐Ala‐OCH3 (7), instead it shows significant loss of t‐butanol. On the other hand, second isomeric pair shows significant loss of t‐butanol and loss of acetone for Boc‐NH‐δ‐Caa‐β‐Caa‐OCH3 (18), whereas these are insignificant for its positional isomer Boc‐NH‐β‐Caa‐δ‐Caa‐OCH3 (13). The tetra‐ and hexapeptide positional isomers also show significant differences in MS2 and MS3 CID spectra. It is observed that ‘b’ ions are abundant when oxazolone structures are formed through five‐membered cyclic transition state and cyclization process for larger ‘b’ ions led to its insignificant abundance. However, b1+ ion is formed in case of δ,α‐dipeptide that may have a six‐membered substituted piperidone ion structure. Furthermore, ESI negative ion MS/MS has also been found to be useful for differentiating these isomeric peptide acids. Thus, the results of MS/MS of pairs of di‐, tetra‐, and hexapeptide positional isomers provide peptide sequencing information and distinguish the positional isomers. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

12.
Methacrylic macromonomers bearing hydrolyzable oligoester segments are prepared by derivatization of oligo(α‐hydroxyalkanoic acids) (obtained by thermal polycondensation) with methacrylic acid and copolymerized with tert‐butyl acrylate.  相似文献   

13.
The Ser, Cys, and His side chains play decisive roles in the syntheses, structures, and functions of proteins and enzymes. For our structural and biomedical investigations of β‐peptides consisting of amino acids with proteinogenic side chains, we needed to have reliable preparative access to the title compounds. The two β3‐homoamino acid derivatives were obtained by Arndt–Eistert methodology from Boc‐His(Ts)‐OH and Fmoc‐Cys(PMB)‐OH (Schemes 2–4), with the side‐chain functional groups' reactivities requiring special precautions. The β2‐homoamino acids were prepared with the help of the chiral oxazolidinone auxiliary DIOZ by diastereoselective aldol additions of suitable Ti‐enolates to formaldehyde (generated in situ from trioxane) and subsequent functional‐group manipulations. These include OH→OtBu etherification (for β2hSer; Schemes 5 and 6), OH→STrt replacement (for β2hCys; Scheme 7), and CH2OH→CH2N3→CH2NH2 transformations (for β2hHis; Schemes 9–11). Including protection/deprotection/re‐protection reactions, it takes up to ten steps to obtain the enantiomerically pure target compounds from commercial precursors. Unsuccessful approaches, pitfalls, and optimization procedures are also discussed. The final products and the intermediate compounds are fully characterized by retention times (tR), melting points, optical rotations, HPLC on chiral columns, IR, 1H‐ and 13C‐NMR spectroscopy, mass spectrometry, elemental analyses, and (in some cases) by X‐ray crystal‐structure analysis.  相似文献   

14.
In the title compound, [Mn(C5H2N2O4)(C12H9N3)2]·H2O, the MnII centre is surrounded by three bidentate chelating ligands, namely, one 6‐oxido‐2‐oxo‐1,2‐dihydropyrimidine‐5‐carboxylate (or uracil‐5‐carboxylate, Huca2−) ligand [Mn—O = 2.136 (2) and 2.156 (3) Å] and two 2‐(2‐pyridyl)‐1H‐benzimidazole (Hpybim) ligands [Mn—N = 2.213 (3)–2.331 (3) Å], and it displays a severely distorted octahedral geometry, with cis angles ranging from 73.05 (10) to 105.77 (10)°. Intermolecular N—H...O hydrogen bonds both between the Hpybim and the Huca2− ligands and between the Huca2− ligands link the molecules into infinite chains. The lattice water molecule acts as a hydrogen‐bond donor to form double O...H—O—H...O hydrogen bonds with the Huca2− O atoms, crosslinking the chains to afford an infinite two‐dimensional sheet; a third hydrogen bond (N—H...O) formed by the water molecule as a hydrogen‐bond acceptor and a Hpybim N atom further links these sheets to yield a three‐dimensional supramolecular framework. Possible partial π–π stacking interactions involving the Hpybim rings are also observed in the crystal structure.  相似文献   

15.
β3‐Peptides consisting of six, seven, and ten homologated proteinogenic amino acid residues have been attached to an α‐heptapeptide (all d‐ amino acid residues; 4 ), to a hexaethylene glycol chain (PEGylation; 5c ), and to dipicolinic acid (DPA derivative 6 ), respectively. The conjugation of the β‐peptides with the second component was carried out through the N‐termini in all three cases. According to NMR analysis (CD3OH solutions), the (M)‐314‐helical structure of the β‐peptidic segments was unscathed in all three chimeric compounds (Figs. 2, 4, and 5). The α‐peptidic section of the α/β‐peptide was unstructured, and so was the oligoethylene glycol chain in the PEGylated compound. Thus, neither does the appendage influence the β‐peptidic secondary structure, nor does the latter cause any order in the attached oligomers to be observed by this method of analysis. A similar conclusion may be drawn from CD spectra (Figs. 1, 3, and 5). These results bode well for the development of delivery systems involving β‐peptides.  相似文献   

16.
The data on temperature, solvent, and high hydrostatic pressure influence on the rate of the ene reactions of 4‐phenyl‐1,2,4‐triazoline‐3,5‐dione ( 1 ) with 2‐carene ( 2 ), and β‐pinene ( 4 ) have been obtained. Ene reactions 1 + 2 and 1 + 4 have high heat effects: ∆Hrn ( 1 + 2 ) −158.4, ∆Hrn( 1 + 4 ) −159.2 kJ mol−1, 25°C, 1,2‐dichloroethane. The comparison of the activation volume (∆V( 1 + 2 ) −29.9 cm3 mol−1, toluene; ∆V( 1 + 4 ) −36.0 cm3 mol−1, ethyl acetate) and reaction volume values (∆Vr‐n( 1 + 2 ) −24.0 cm3 mol−1, toluene; ∆Vr‐n( 1 + 4 ) −30.4 cm3 mol−1, ethyl acetate) reveals more compact cyclic transition states in comparison with the acyclic reaction products 3 and 5 . In the series of nine solvents, the reaction rate of 1+2 increases 260‐fold and 1+4 increases 200‐fold, respectively, but not due to the solvent polarity.  相似文献   

17.
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   

18.
The NCAs of the following five amino acids were polymerized in bulk at 120 °C without addition of a catalyst or initiator: sarcosine (Sar), L ‐alanine (L ‐Ala), D ,L ‐phenylalanine (D ,L ‐Phe), D ,L ‐leucine (D ,L ‐Leu) and D ,L ‐valine (D,L ‐Val). The virgin reaction products were characterized by viscosity measurements 13C NMR spectroscopy and MALDI‐TOF mass spectrometry. In addition to numerous low molar mass byproducts cyclic polypeptides were formed as the main reaction products in the mass range above 800 Da. Two types of cyclic oligo‐ and polypeptides were detected in all cases with exception of sarcosine NCA, which only yielded one class of cyclic polypeptides. The efficient formation of cyclic oligo‐ and polypeptides explains why high molar mass polymers cannot be obtained by thermal polymerizations of α‐amino acid NCAs. Various polymerization mechanisms were discussed. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 4012–4020, 2008  相似文献   

19.
20.
The lithium salt (HC–NCMe3)2SiFNLiR ( 1 ) R = C6H3(2,6‐CHMe2)2 reacts with trichlorogallium under displacement of the lithium ion by GaCl3 to give the adduct [(HC–NCMe3)2SiFN] [(GaCl3)R·Li(thf)4]+ ( 1 ). Compound 1 thermally loses LiCl and forms the bicyclic ring intermediates V and VI . Compound  VI adds the aniline H2NC6H3(2,6‐CHMe2)2 and the unsaturated, seven‐membered ring compound –NCMe3–CH2–CH=NCMe3GaCl2–NR–SiFNHR– ( 2 ) is obtained. The addition is accompanied by an enamine‐imine‐tautomerism and proves the Lewis acid character of the silicon atom in an unknown 3‐center‐2‐electron interaction of one nitrogen atom with the silicon and gallium atoms. Quantum chemical calculations of the thermal isomerisation process and crystal structures of 1 and 2 are reported.  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