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 a general and operationally simple method for the solid phase synthesis of α‐ketoamide peptides using standard Fmoc solid phase peptide synthesis. The method delivers deprotected peptide α‐ketoamides directly upon resin cleavage without any additional steps, and tolerates all side chain functional groups. A small collection of C‐terminal and internal α‐ketoamide peptides – including two reported protease inhibitors (HCV and SUB1) – were prepared in good yields. In addition, we demonstrate that our method serves as versatile platform for the convenient preparation of cyclic α‐ketoamide peptides, photocagged peptide α‐ketoamides, and fluorescently labeled peptides.  相似文献   

2.
We report non‐chiral amino acid residues cis‐ and trans‐1,4‐diaminocyclohexane‐1‐carboxylic acid (cyclo‐ornithine, cO) that exhibit unprecedented stereospecific control of backbone dissociations of singly charged peptide cations and hydrogen‐rich cation radicals produced by electron‐transfer dissociation. Upon collision‐induced dissociation (CID) in the slow heating regime, peptide cations containing trans‐cO residues undergo facile backbone cleavages of amide bonds C‐terminal to trans‐cO. By contrast, peptides with cis‐cO residues undergo dissociations at several amide bonds along the peptide ion backbone. Diastereoisomeric cO‐containing peptides thus provide remarkably distinct tandem mass spectra. The stereospecific effect in CID of the trans‐cO residue is explained by syn‐facially directed proton transfer from the 4‐ammonium group at cO to the C‐terminal amide followed by neighboring group participation in the cleavage of the CO―NH bond, analogous to the aspartic acid and ornithine effects. Backbone dissociations of diastereoisomeric cO‐containing peptide ions generate distinct [bn]+‐type fragment ions that were characterized by CID‐MS3 spectra. Stereospecific control is also reported for electron‐transfer dissociation of cis‐ and trans‐cO containing doubly charged peptide ions. The stereospecific effect upon electron transfer is related to the different conformations of doubly charged peptide ions that affect the electron attachment sites and ensuing N―Cα bond dissociations.  相似文献   

3.
Glycine (Gly) is incorporated in roughly half of all known peptaibiotic (nonribosomally biosynthesized antibiotic peptides of fungal origin) sequences and is the residue with the greatest conformational flexibility. The conformational space of Aib (α‐aminoisobutyric acid) is severely restricted by the second methyl group attached to the Cα atom. Most of the crystal structures containing Aib are N‐terminal protected. Deprotection of the N‐ or C‐terminus of peptides may alter the hydrogen‐bonding scheme and/or the structure and may facilitate crystallization. The structure reported here for glycyl‐α‐aminoisobutyrylglycyl‐α‐aminoisobutyric acid tert‐butyl ester, C16H30N4O5, describes the first N‐terminal‐unprotected (Gly‐Aib)n peptide. The achiral peptide could form an intramolecular hydrogen bond between the C=O group of Gly1 and the N—H group of Aib4. This hydrogen bond is found in all tetrapeptides and N‐terminal‐protected tripeptides containing Aib, apart from one exception. In the present work, this hydrogen bond is not observed (N...O = 5.88 Å). Instead, every molecule is hydrogen bonded to six other symmetry‐related molecules with a total of eight hydrogen bonds per molecule. The backbone conformation starts in the right‐handed helical region (and the left‐handed helical region for the inverted molecule) and reverses the screw sense in the last two residues.  相似文献   

4.
Hydrogens connected to α‐carbon (α‐C) of amino acid residues are usually resistant to hydrogen‐deuterium exchange (HDX) unless reaction conditions promote racemization. Although N‐methylglycine (sarcosine) residue has been found in biologically active peptide such as cyclosporine, to the best of our knowledge, the HDX of α‐C protons of this residue was not explored yet. Here, we presented a new and efficient methodology of α‐C deuteration in sarcosine residues under basic aqueous conditions. The deuterons, introduced at α‐C atom, do not undergo back‐exchange in acidic aqueous solution. The electrospray ionization‐MS and MS/MS experiments on proposed model peptides confirmed the HDX at α‐C and revealed the unexpected hydrogen scrambling in sarcosine‐containing peptides. Although the observed HDX of α‐C protons is only successful in N‐acylglycine when the amide possesses a certain degree of alkylation, it offers a new approach to the analysis of sarcosine‐containing peptides such as cyclosporine. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

5.
The gas‐phase free radical initiated peptide sequencing (FRIPS) fragmentation behavior of o‐TEMPO‐Bz‐conjugated peptides with an intra‐ and intermolecular disulfide bond was investigated using MSn tandem mass spectrometry experiments. Investigated peptides included four peptides with an intramolecular cyclic disulfide bond, Bactenecin (RLC RIVVIRVC R), TGF‐α (C HSGYVGVRC ), MCH (DFDMLRC MLGRVFRPC WQY) and Adrenomedullin (16–31) (C RFGTC TVQKLAHQIY), and two peptides with an intermolecular disulfide bond. Collisional activation of the benzyl radical conjugated peptide cation, which was generated through the release of a TEMPO radical from o‐TEMPO‐Bz‐conjugated peptides upon initial collisional activation, produced a large number of peptide backbone fragments in which the S? S or C? S bond was readily cleaved. The observed peptide backbone fragments included a‐, c‐, x‐ or z‐types, which indicates that the radical‐driven peptide fragmentation mechanism plays an important role in TEMPO‐FRIPS mass spectrometry. FRIPS application of the linearly linked disulfide peptides further showed that the S? S or C? S bond was selectively and preferentially cleaved, followed by peptide backbone dissociations. In the FRIPS mass spectra, the loss of ?SH or ?SSH was also abundantly found. On the basis of these findings, FRIPS fragmentation pathways for peptides with a disulfide bond are proposed. For the cleavage of the S? S bond, the abstraction of a hydrogen atom at Cβ by the benzyl radical is proposed to be the initial radical abstraction/transfer reaction. On the other hand, H‐abstraction at Cα is suggested to lead to C? S bond cleavage, which yields [ion ± S] fragments or the loss of ?SH or ?SSH. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

6.
Matrix‐assisted laser desorption/ionization in‐source decay (MALDI‐ISD) is initiated by hydrogen transfer from matrix molecules to the carbonyl oxygen of peptide backbone with subsequent radical‐induced cleavage leading to c′/z? fragments pair. MALDI‐ISD is a very powerful method to obtain long sequence tags from proteins or to do de novo sequencing of peptides. Besides classical fragmentation, MALDI‐ISD also shows specific fragments for which the mechanism of formation enlightened the MALDI‐ISD process. In this study, the MALDI‐ISD mechanism is reviewed, and a specific mechanism is studied in details: the N‐terminal side of Cys residue (Xxx‐Cys) is described to promote the generation of c′ and w fragments in MALDI‐ISD. Our data suggest that for sequences containing Xxx‐Cys motifs, the N–Cα bond cleavage occurs following the hydrogen attachment to the thiol group of Cys side‐chain. The c?/w fragments pair is formed by side‐chain loss of the Cys residue with subsequent radical‐induced cleavage at the N–Cα bond located at the left side (N‐terminal direction) of the Cys residue. This fragmentation pathway preferentially occurs at free Cys residue and is suppressed when the cysteines are involved in disulfide bonds. Hydrogen attachment to alkylated Cys residues using iodoacetamide gives free Cys residue by the loss of ?CH2CONH2 radical. The presence of alkylated Cys residue also suppress the formation of c?/w fragments pair via the (Cβ)‐centered radical, whereas w fragment is still observed as intense signal. In this case, the z? fragment formed by hydrogen attachment of carbonyl oxygen followed side‐chain loss at alkylated Cys leads to a w fragment. Hydrogen attachment on peptide backbone and side‐chain of Cys residue occurs therefore competitively during MALDI‐ISD process. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

7.
The title dipeptide, 1‐(tert‐butoxy­carbonyl‐d ‐alanyl)‐N‐iso­propyl‐l ‐pipecol­amide or Boc‐d ‐Ala‐l ‐Pip‐NHiPr (H‐Pip‐OH is pipecolic acid or piperidine‐2‐carboxylic acid), C17H31N3­O4, with a d –l heterochiral sequence, adopts a type II′β‐­turn conformation, with all‐trans amide functions, where the C‐terminal amide NH group interacts with the Boc carbonyl O atom to form a classical i+3 i intramolecular hydrogen bond. The Cα substituent takes an axial position [Hα (Pip) equatorial] and the trans pipecolamide function is nearly planar.  相似文献   

8.
《Electroanalysis》2017,29(7):1700-1711
A selective method based on derivatization with 2,4‐dinitrophenylhydrazine (DNPH) is described for the determination of several carbonyl compounds. The factors affecting the derivatization reaction of aldehyde and DNPH were investigated. The product of the derivatization reaction has been characterized by UV/Vis spectrophotometry, NMR, infrared spectroscopy and cyclic voltammetry. Then, an electrochemical study for the determination of aldehyde‐2.4‐dinitrophenylhydrazone was performed at glassy carbon electrode (GCE) using square wave voltammetry (SWV). After the optimization of experimental parameters, the limits of detection (at 3σ ) obtained for all aldehyde‐2,4‐DNPH were varied from 15.82 to 78.39 μmol L−1 and relative standard deviations were between 1.8 and 4.5%. Finally, the proposed method was applied to determine the aldehydes concentration in drinking water and orange juice samples with satisfactory results.  相似文献   

9.
To develop the new‐type poly(L ‐lactide)‐based biomedical material having a wettable surface, the synthesis of poly(L ‐lactide) with one terminal D ‐glucose residue was investigated. After the hydroxyl group at 1‐C of α‐tetrabenzyl glucose, α‐Glc(Bzl)4, was converted to the corresponding potassium alkoxide by using potassium tert‐butoxide, L ‐lactide (L ‐LA) was polymerized in the presence of α‐Glc(Bzl)4‐OK as an initiator in tetrahydrofuran at room temperature to prepare α‐Glc(Bzl)4‐polyLA. Subsequently, the removal of O‐protecting benzyl groups in the terminal α‐Glc(Bzl)4 residue was carried out by hydrogenolysis with Pd/C to obtain the objective D ‐glucose‐end‐capped polyLA, α‐Glc‐polyLA. The wettability of surface of the α‐Glc‐polyLA material is discussed using the difference of the dynamic contact angle between a α‐Glc‐polyLA/homopolyLA blend film and a film of the polyLA homopolymer.  相似文献   

10.
An α‐helical model peptide (Ac‐EAEKAAKE‐X‐EKAAKEAEK‐amide) was used as a template to examine the efficacy of conventional reversed‐phase high‐performance liquid chromatography (RP‐HPLC) in separating peptide analogs with single substitutions (at position X) of diasteromeric amino acids Ile, allo‐Ile, d ‐Ile and d ‐allo‐Ile. We compared differences in peptide retention behavior on a C8 column and a C18 column at different temperatures. We demonstrated how subtle differences in peptide secondary structure affected by the different substitutions of amino acids with identical overall hydrophobicity enabled effective resolution of these peptide analogs. We also demonstrated the ability of RP‐HPLC to separate Ile‐ and allo‐Ile‐substituted analogs of a 26‐residue α‐helical antimicrobial peptide (AMP), with the substitution site towards the C‐terminus of the α‐helix. These peptides show different values of antibacterial activity and hemolytic activity, and different selectivity against bacteria and human cells. Our results underline the ability of RP‐HPLC to resolve even difficult diasteromeric peptide mixtures as well as its value in monitoring very subtle hydrophobicity changes in de novo‐designed AMP. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

11.
The importance of β‐peptides lies in their ability to mimic the conformational behavior of α‐peptides, even with a much shorter chain length, and in their resistance to proteases. To investigate the effect of substitution of β‐peptides on their dominant fold, we have carried out a molecular‐dynamics (MD) simulation study of two tetrapeptides, Ac‐(2R,3S)‐β2,3hVal(αMe)‐(2S)‐β2hPhe‐(R)‐β3hLys‐(2R,3S)‐β2,3‐Ala(αMe)‐NH2, differing in the substitution at the Cα of Phe2 (pepF with F, and pepH with H). Three simulations, unrestrained (UNRES), using 3J‐coupling biasing with local elevation in combination with either instantaneous (INS) or time‐averaging (AVE) NOE distance restraining, were carried out for each peptide. In the unrestrained simulations, we find three (pepF) and two (pepH) NOE distance bound violations of maximally 0.22 nm that involve the terminal residues. The restrained simulations match both the NOE distance bounds and 3J‐values derived from experiment. The fluorinated peptide shows a slightly larger conformational variability than the non‐fluorinated one.  相似文献   

12.
The convergent synthesis of proteins by multiple ligations requires segments protected at the N‐ and/or C‐terminus with masking groups that are orthogonal to the acid‐ and base‐labile protecting groups used in Fmoc‐SPPS. They must be stable to solid‐phase peptide synthesis, HPLC purification, and ligation conditions and easily removed in the presence of unprotected side chains. In this report, we document photolabile protecting groups for both α‐ketoacids and hydroxylamines, the key functional groups employed in the α‐ketoacid–hydroxylamine (KAHA) ligation. The novel photoprotected α‐ketoacid is easily installed onto numerous different C‐terminal peptide α‐ketoacids and removed by UV light under aqueous conditions. These advances were applied to the one‐pot synthesis of NEDD8, an important modifier protein, by three different convergent routes. These new protecting groups provide greater flexibility on the order of fragment assembly and reduce the number of reaction and purification steps needed for protein synthesis with the KAHA ligation.  相似文献   

13.
Nα[(tert-Butoxy)carbonyl]-2-diazo-L -histidine methyl ester 1 was synthesized starting from the corresponding L-histidine derivative. The physico-chemical properties of this new photoactivatable amino-acid derivative were established. The synthetic precursor of 1 , 2-amino-L -histidine derivative 3 , was best isolated and characterized as 2-amino-Nα-[(tert-butoxy)carbonyl]-Nτ-tosyl-L -histidine methyl ester ( 4 ). Selective deprotections of 4 (Nα-Boc, Nα-Tos, COOMe) were achieved, thus allowing the use of the corresponding products in peptide synthesis. The optically active dipeptides 8 and 9 were synthesized by coupling 2-amino-Nτ-tosyl-L -histidine methyl ester ( 5 ) with N-[(tert-butoxy)carbonyl]-L -alanine and Nα-[(tert-butoxy)carbonyl]-Nτ-tosyl-L -histidine ( 6 ) with L-alanine methyl ester, respectively. The question of selective diazotization of a 2-aminohistidine residue in a synthetic peptide was studied using competitive diazotizations between 2-amino-1H-imidazole and several amino-acid derivatives susceptible to undergo nitrosylation. The results show that synthetic photoactivatable peptides incorporating a 2-diazohistidine residue might become useful photoaffinity probes.  相似文献   

14.
A systematic conformational analysis on blocked β‐amino acids as constituents of β‐peptides by ab initio MO theory reveals that the conformer pool of β‐peptide monomers is essentially determined by the conformation of simple submonomer fragments. The influence of single and multiple substitutions at the C(α) and C(β) backbone atoms on the intrinsic folding properties of the monomers was estimated both in the single‐molecule approximation and in a polar solvent continuum, applying a quantum‐chemical SCRF model. Substitution at C(β) has a higher impact on the β‐amino acid conformation than a substitution at C(α). It can be shown that the conformations of important periodic secondary structures in β‐peptides belong to the conformer pool of the monomers, even for those secondary‐structure elements where H‐bond formation appears only in longer sequences. Rules for design of special secondary‐structure types by selection of an actual substituent pattern in the β‐amino acid constituents have been derived within the monomer approach.  相似文献   

15.
The fragment β(25–35) of the amyloid β‐peptide, like its parent βA4, has shown neurotrophic and late neurotoxic activities in cultured cells. The 3D structure of this important peptide was examined by 1H and 13C 2D‐NMR and MD simulations in DMSO‐d6 and water. The NMR parameters of chemical shift, 3J(N,Hα) coupling constants, temperature coefficients of NH chemical shifts and the pattern of intra and inter‐residue NOEs were used to deduce the structures. In DMSO‐d6, the peptide was found to take up a type I β‐turn around the C‐terminal residues Ile8–Gly9–Leu10–Met11, whereas in water at pH 5.5, it adopts a random coil conformation. This is only the second report of a β‐turn in the β‐amyloid class of peptides. The solution structures generated using restrained molecular dynamics were refined by MARDIGRAS to an R factor of 0.33 in the case of DMSO‐d6 and to 0.56 for water. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

16.
Five β‐peptide thioesters ( 1 – 5 , containing 3, 4, 10 residues) were prepared by manual solid‐phase synthesis and purified by reverse‐phase preparative HPLC. A β‐undecapeptide ( 6 ) and an α‐undecapeptide ( 7 ) with N‐terminal β3‐HCys and Cys residues were prepared by manual and machine synthesis, respectively. Coupling of the thioesters with the cysteine derivatives in the presence of PhSH (Scheme and Fig. 1) in aqueous solution occurred smoothly and quantitatively. Pentadeca‐ and heneicosapeptides ( 8 – 10 ) were isolated, after preparative RP‐HPLC purification, in yields of up to 60%. Thus, the so‐called native chemical ligation works well with β‐peptides, producing larger β3‐ and α/β3‐mixed peptides. Compounds 1 – 10 were characterized by high‐resolution mass spectrometry (HR‐MS) and by CD spectroscopy, including temperature and concentration dependence. β‐Peptide 9 with 21 residues shows an intense negative Cotton effect near 210 nm but no zero‐crossing above 190 nm, (Figs. 2–4), which is characteristic of β‐peptidic 314‐helical structures. Comparison of the CD spectra of the mixed α/β‐pentadecapeptide ( 10 ) and a helical α‐peptide (Fig. 5) indicate the presence of an α‐peptidic 3.613 helix.  相似文献   

17.
For the first time, the Petasis (borono‐Mannich) reaction is employed for the multicomponent labeling and stapling of peptides. The report includes the solid‐phase derivatization of peptides at the N‐terminus, Lys, and N?‐MeLys side‐chains by an on‐resin Petasis reaction with variation of the carbonyl and boronic acid components. Peptides were simultaneously functionalized with aryl/vinyl substituents bearing fluorescent/affinity tags and oxo components such as dihydroxyacetone, glyceraldehyde, glyoxylic acid, and aldoses, thus encompassing a powerful complexity‐generating approach without changing the charge of the peptides. The multicomponent stapling was conducted in solution by linking N?‐MeLys or Orn side‐chains, positioned at i, i+7 and i, i+4, with aryl tethers, while hydroxy carbonyl moieties were introduced as exocyclic fragments. The good efficiency and diversity oriented character of these methods show prospects for peptide drug discovery and chemical biology.  相似文献   

18.
For the first time, the Petasis (borono‐Mannich) reaction is employed for the multicomponent labeling and stapling of peptides. The report includes the solid‐phase derivatization of peptides at the N‐terminus, Lys, and N?‐MeLys side‐chains by an on‐resin Petasis reaction with variation of the carbonyl and boronic acid components. Peptides were simultaneously functionalized with aryl/vinyl substituents bearing fluorescent/affinity tags and oxo components such as dihydroxyacetone, glyceraldehyde, glyoxylic acid, and aldoses, thus encompassing a powerful complexity‐generating approach without changing the charge of the peptides. The multicomponent stapling was conducted in solution by linking N?‐MeLys or Orn side‐chains, positioned at i, i+7 and i, i+4, with aryl tethers, while hydroxy carbonyl moieties were introduced as exocyclic fragments. The good efficiency and diversity oriented character of these methods show prospects for peptide drug discovery and chemical biology.  相似文献   

19.
We present a molecular‐dynamics simulation study of an α‐heptapeptide containing an α‐aminoisobutyric acid (=2‐methylalanine; Aib) residue, Val1‐Ala2‐Leu3‐Aib4‐Ile5‐Met6‐Phe7, and a quantum‐mechanical (QM) study of simplified models to investigate the propensity of the Aib residue to induce 310/α‐helical conformation. For comparison, we have also performed simulations of three analogues of the peptide with the Aib residue being replaced by L ‐Ala, D ‐Ala, and Gly, respectively, which provide information on the subtitution effect at C(α) (two Me groups for Aib, one for L ‐Ala and D ‐Ala, and zero for Gly). Our simulations suggest that, in MeOH, the heptapeptide hardly folds into canonical helical conformations, but appears to populate multiple conformations, i.e., C7 and 310‐helical ones, which is in agreement with results from the QM calculations and NMR experiments. The populations of these conformations depend on the polarity of the solvent. Our study confirms that a short peptide, though with the presence of an Aib residue in the middle of the chain, does not have to fold to an α‐helical secondary structure. To generate a helical conformation for a linear peptide, several Aib residues should be present in the peptide, either sequentially or alternatively, to enhance the propensity of Aib‐containing peptides towards the helical conformation. A correction of a few of the published NMR data is reported.  相似文献   

20.
The prominence of endogenous peptide ligands targeted to receptors makes peptides with the desired binding activity good molecular scaffolds for drug development. Minor modifications to a peptide's primary sequence can significantly alter its binding properties with a receptor, and screening collections of peptide mutants is a useful technique for probing the receptor–ligand binding domain. Unfortunately, the combinatorial growth of such collections can limit the number of mutations which can be explored using structure‐based molecular docking techniques. Genetic algorithm managed peptide mutant screening (GAMPMS) uses a genetic algorithm to conduct a heuristic search of the peptide's mutation space for peptides with optimal binding activity, significantly reducing the computational requirements of the virtual screening. The GAMPMS procedure was implemented and used to explore the binding domain of the nicotinic acetylcholine receptor (nAChR) ‐isoform with a library of 64,000 α‐conotoxin (α‐CTx) MII peptide mutants. To assess GAMPMS's performance, it was compared with a virtual screening procedure that used AutoDock to predict the binding affinity of each of the α‐CTx MII peptide mutants with the ‐nAChR. The GAMPMS implementation performed AutoDock simulations for as few as 1140 of the 64,000 α‐CTx MII peptide mutants and could consistently identify a set of 10 peptides with an aggregated binding energy that was at least 98% of the aggregated binding energy of the 10 top peptides from the exhaustive AutoDock screening. © 2015 Wiley Periodicals, Inc.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号