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 315 毫秒
1.
The Pictet–Spengler (PS) cyclizations of β3hTrp derivatives as arylethylamine substrates were performed with L‐α‐amino and D‐α‐amino aldehydes as carbonyl components. During the PS reaction, a new stereogenic center was created, and the mixture of cis/trans 1,3‐disubstituted 1,2,3,4‐tetrahydro‐β‐carbolines was obtained. The ratio of cis/trans diastereomers depends on the stereogenic centre of used amino aldehyde and the size of substituents. It was confirmed by 1H and 2D NMR (ROESY) spectra. The conformations of cyclic products were studied by 2D NMR ROESY spectra. Products of the PS condensation after removal of protecting group(s) can be incorporated into a peptide chain as tryptophan mimetics with the possibility of the β‐turn induction. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
Dipeptides with an aromatic residue at the N–terminal position induced lower inward currents or blocked leak currents in Xenopus oocytes expressing the proton-coupled peptide transporter AtPTR1 or AtPTR5 of Arabidopsis thaliana compared with dipeptides with an aromatic residue at the C–terminal position. Here, AtPTR1 and AtPTR5 were expressed in a yeast mutant of peptide transporter (ptr2) with tryptophan auxotrophy. Growth assays showed that Trp–Ala could be transported by both AtPTR1 and AtPTR5 as efficiently as Ala–Trp. Our data suggested that the previous finding in Xenopus oocytes might be an artifact of heterologous expression, and that AtPTR1 and AtPTR5 expressed in yeast could transport dipeptides with an aromatic residue at the N–terminal position.  相似文献   

3.
The importance of dipeptides both in medicinal and pharmacological fields is well documented and many efforts have been made to find simple and efficient methods for their synthesis. For this reason, we have investigated the synthesis of α‐N‐protected dipeptide acids by reacting the easily accessible mixed anhydride of α‐N‐protected amino acids with free amino acids under different reaction conditions. The combination of TBA‐OH and DMSO has been found to be the best to overcome the low solubility of amino acids in organic solvents. Under these experimental conditions, the homogeneous phase condensation reaction occurs rapidly and without detectable epimerization. The present method is also applicable to side‐chain unprotected Tyr, Trp, Glu, and Asp but not Lys. This latter residue is able to engage two molecules of mixed anhydride giving the corresponding isotripeptide. Moreover, the applicability of this protocol for the synthesis of tri‐ and tetrapeptides has been tested. This approach reduces the need for protecting groups, is cost effective, scalable, and yields dipeptide acids that can be used as building blocks in the synthesis of larger peptides. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
An aldo‐keto reductase AKR5C3 from Gluconobacter oxydans (designated as Gox0644) is a useful enzyme with various substrates, including aldehydes, diacetyl, keto esters, and α‐ketocarbonyl compounds. The crystal structures of AKR5C3 in apoform in complex with NADPH and the D53A mutant (AKR5C3‐D53A) in complex with NADPH are presented herein. Structure comparison and site‐directed mutagenesis combined with biochemical kinetics analysis reveal that the conserved Asp53 in the AKR5C3 catalytic tetrad has a crucial role in securing active pocket conformation. The gain‐of‐function Asp53 to Ala mutation triggers conformational changes on the Trp30 and Trp191 side chains, improving NADPH affinity to AKR5C3, which helps increase catalytic efficiency. The highly conserved Trp30 and Trp191 residues interact with the nicotinamide moiety of NADPH and help form the NADPH‐binding pocket. The AKR5C3‐W30A and AKR5C3‐W191Y mutants show decreased activities, confirming that both residues facilitate catalysis. Residue Trp191 is in the loop structure, and the AKR5C3‐W191Y mutant does not react with benzaldehyde, which might also determine substrate recognition. Arg192, which is involved in the substrate binding, is another important residue. The introduction of R192G increases substrate‐binding affinity by improving hydrophobicity in the substrate‐binding pocket. These results not only supplement the AKRs superfamily with crystal structures but also provide useful information for understanding the catalytic properties of AKR5C3 and guiding further engineering of this enzyme.  相似文献   

5.
Neisseria meningitidis is the causative agent of meningitis and meningococcal septicemia is a major cause of disease worldwide, resulting in brain damage and hearing loss, and can be fatal in a large proportion of cases. The enzyme 3‐deoxy‐d ‐arabino‐heptulosonate 7‐phosphate synthase (DAH7PS) catalyzes the first reaction in the shikimate pathway leading to the biosynthesis of aromatic metabolites including the aromatic acids l ‐Trp, l ‐Phe, and l ‐Tyr. This pathway is absent in humans, meaning that enzymes of the pathway are considered as potential candidates for therapeutic intervention. As the entry point, feedback inhibition of DAH7PS by pathway end products is a key mechanism for the control of pathway flux. The structure of the single DAH7PS expressed by N. meningitidis was determined at 2.0 Å resolution. In contrast to the other DAH7PS enzymes, which are inhibited only by a single aromatic amino acid, the N. meningitidis DAH7PS was inhibited by all three aromatic amino acids, showing greatest sensitivity to l ‐Phe. An N. meningitidis enzyme variant, in which a single Ser residue at the bottom of the inhibitor‐binding cavity was substituted to Gly, altered inhibitor specificity from l ‐Phe to l ‐Tyr. Comparison of the crystal structures of both unbound and Tyr‐bound forms and the small angle X‐ray scattering profiles reveal that N. meningtidis DAH7PS undergoes no significant conformational change on inhibitor binding. These observations are consistent with an allosteric response arising from changes in protein motion rather than conformation, and suggest ligands that modulate protein dynamics may be effective inhibitors of this enzyme.  相似文献   

6.
Imidazole dipeptides, such as carnosine (β‐alanyl‐l ‐histidine) and anserine (β‐alanyl‐Nπ‐methyl‐l ‐histidine), are highly localized in excitable tissues, including skeletal muscle and nervous tissue, and play important roles such as scavenging reactive oxygen species and quenching reactive aldehydes. We have demonstrated several reactions between imidazole dipeptides (namely, carnosine, and anserine) and a lipid peroxide‐derived reactive aldehyde 4‐oxo‐2(E)‐nonenal. Seven carnosine adducts and two anserine adducts were characterized using liquid chromatography/electrospray ionization‐multiple‐stage mass spectrometry. Adduct formation occurred between imidazole dipeptides and 4‐oxo‐2(E)‐nonenal mainly through Michael addition, Schiff base formation, and/or Paal‐Knorr reaction. The reactions were much more complicated than the reaction with a similar lipid peroxide‐derived reactive aldehyde, 4‐hydroxy‐2(E)‐nonenal.  相似文献   

7.
BacD is an ATP‐dependent dipeptide ligase responsible for the biosynthesis of L ‐alanyl‐L ‐anticapsin, a precursor of an antibiotic produced by Bacillus spp. In contrast to the well‐studied and phylogenetically related D ‐alanine: D ‐alanine ligase (Ddl), BacD synthesizes dipeptides using L ‐amino acids as substrates and has a low substrate specificity in vitro. The enzyme is of great interest because of its potential application in industrial protein engineering for the environmentally friendly biological production of useful peptide compounds, such as physiologically active peptides, artificial sweeteners and antibiotics, but the determinants of its substrate specificity and its catalytic mechanism have not yet been established due to a lack of structural information. In this study, we report the crystal structure of BacD in complex with ADP and an intermediate analog, phosphorylated phosphinate L ‐alanyl‐L ‐phenylalanine, refined to 2.5‐Å resolution. The complex structure reveals that ADP and two magnesium ions bind in a manner similar to that of Ddl. However, the dipeptide orientation is reversed, and, concomitantly, the entrance to the amino acid binding cavity differs in position. Enzymatic characterization of two mutants, Y265F and S185A, demonstrates that these conserved residues are not catalytic residues at least in the reaction where L ‐phenylalanine is used as a substrate. On the basis of the biochemical and the structural data, we propose a reaction scheme and a catalytic mechanism for BacD.  相似文献   

8.
The reaction of various dipeptides with glyoxal at 100°C, pH 5.0 was studied. Carbon dioxide, ammonia, amino acids and aldehydes were detected from the reaction solutions. Besides, a series of new pyrazinones—2-(3′-alkyl-2′-oxo-pyrazin-1′-yl)alkyl acid—were isolated, and their chemical structures were confirmed by UV, IR, MS and NMR spectra. These pyrazinones seem to play a role in the browning of the reaction.

In the reaction with glyoxal, acetaldehyde and glucose, reactivity of peptides was proved to be much higher than that of amino acids.

The reaction mechanism of dipeptides with glyoxal was also proposed.  相似文献   

9.
N‐(tert‐butyloxycarbonyl) or N‐(9‐fluorenylmethoxycarbonyl) dipeptides with C‐terminal (Z)‐α,β‐didehydrophenylalanine (?ZPhe), (Z)‐α,β‐didehydrotyrosine (?ZTyr), (Z)‐α,β‐didehydrotryptophan (?ZTrp), (Z)‐α,β‐didehydromethionine (?ZMet), (Z)‐α,β‐didehydroleucine (?ZLeu), and (Z/E)‐α,β‐didehydroisoleucine (?Z/EIle) were synthesised from their saturated analogues via oxidation of intermediate 2,5‐disubstituted‐oxazol‐5‐(4H)‐ones (also known as azlactones) with pyridinium tribromide followed by opening of the produced unsaturated oxazol‐5‐(4H)‐one derivatives in organic‐aqueous solution with a catalytic amount of trifluoroacetic acid or by a basic hydrolysis. In all cases, a very strong preference for Z isomers of α,β‐didehydro‐α‐amino acid residues was observed except of the ΔIle, which was obtained as the equimolar mixture of Z and E isomers. Reasons for the (Z)‐stereoselectivity and the increased stability of the aromatic α,β‐didehydro‐α‐amino acid residue oxazol‐5‐(4H)‐ones over the corresponding aliphatic ones are also discussed. It is the first use of such a procedure to synthesise peptides with the C‐terminal unsaturated residues and a peptide with 2 consecutive ΔPhe residues. This approach is very effective especially in the synthesis of peptides with aliphatic α,β‐didehydro‐α‐amino acid residues that are difficult to obtain by other methods. It allowed the first synthesis of the ?Met residue. It is also more cost‐effective and less laborious than other synthesis protocols. The dipeptide building blocks obtained were used in the solid‐phase synthesis of model peptides on a polystyrene‐based solid support. Peptides containing aromatic α,β‐didehydro‐α‐amino acid residues were obtained with PyBOP or TBTU as a coupling agent with good yields and purities. In the case of aliphatic α,β‐didehydro‐α‐amino acid residues, a good efficiency was achieved only with DPPA as a coupling agent.  相似文献   

10.
The direct asymmetric aldol reactions of equivalent molar amounts of aldehydes and ketones were carried out at −20 °C over alkaline Al2O3 with 20 mol % of Pro‐Trp as catalyst and 20 mol % of N‐methylmorpholine or 1,4‐diazabicyclo[2.2.2]octane as additive. After simple and environmentally friendly work‐up, moderate to high isolated yields (up to 95%), good diastereoselectivities (>99:1), and enantioselectivities (up to 98% ee) have been achieved for the reactions of different kinds of ketones with various aldehydes. The catalytic system could be reused without decrease of activity by addition of 10 mol % catalyst and base in the catalytic system. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

11.
Glutamic acid is an abundant amino acid that lends a characteristic umami taste to foods. In fermented foods, glutamic acid can be found as a free amino acid formed by proteolysis or as a non‐proteolytic derivative formed by microorganisms. The aim of the present study was to identify different structures of glutamic acid derivatives in a typical fermented protein‐based food product, soy sauce. An acidic fraction was prepared with anion‐exchange solid‐phase extraction (SPE) and analyzed by UPLC/MS/MS and UPLC/TOF‐MS. α‐Glutamyl, γ‐glutamyl, and pyroglutamyl dipeptides, as well as lactoyl amino acids, were identified in the acidic fraction of soy sauce. They were chemically synthesized for confirmation of their occurrence and quantified in the selected reaction monitoring (SRM) mode. Pyroglutamyl dipeptides accounted for 770 mg/kg of soy sauce, followed by lactoyl amino acids (135 mg/kg) and γ‐glutamyl dipeptides (70 mg/kg). In addition, N‐succinoylglutamic acid was identified for the first time in food as a minor compound in soy sauce (5 mg/kg).  相似文献   

12.
The Type IX secretion system (T9SS) is present in over 1000 sequenced species/strains of the Fibrobacteres‐Chlorobi‐Bacteroidetes superphylum. Proteins secreted by the T9SS have an N‐terminal signal peptide for translocation across the inner membrane via the SEC translocon and a C‐terminal signal for secretion across the outer membrane via the T9SS. Nineteen protein components of the T9SS have been identified including three, SigP, PorX and PorY that are involved in regulation. The inner membrane proteins PorL and PorM and the outer membrane proteins PorK and PorN interact and a complex comprising PorK and PorN forms a large ring structure of 50 nm in diameter. PorU, PorV, PorQ and PorZ form an attachment complex on the cell surface of the oral pathogen, Porphyromonas gingivalis. P. gingivalis T9SS substrates bind to PorV suggesting that after translocation PorV functions as a shuttle protein to deliver T9SS substrates to the attachment complex. The PorU component of the attachment complex is a novel Gram negative sortase which catalyses the cleavage of the C‐terminal signal and conjugation of the protein substrates to lipopolysaccharide, anchoring them to the cell surface. This review presents an overview of the T9SS focusing on the function of T9SS substrates and machinery components.  相似文献   

13.
D ‐Tetronitrose is a nitro‐containing tetradeoxysugar found attached to the antitumor and antibacterial agent tetrocarcin A. The biosynthesis of this highly unusual sugar in Micromonospora chalcea requires 10 enzymes. The fifth step in the pathway involves the transfer of a methyl group from S‐adenosyl‐L ‐methionine (SAM) to the C‐3′ carbon of dTDP‐3‐amino‐2,3,6‐trideoxy‐4‐keto‐D ‐glucose. The enzyme responsible for this transformation is referred to as TcaB9. It is a monomeric enzyme with a molecular architecture based around three domains. The N‐terminal motif contains a binding site for a structural zinc ion. The middle‐ and C‐terminal domains serve to anchor the SAM and dTDP–sugar ligands, respectively, to the protein, and the active site of TcaB9 is wedged between these two regions. For this investigation, the roles of Tyr 76, His 181, Tyr 222, Glu 224, and His 225, which form the active site of TcaB9, were probed by site‐directed mutagenesis, kinetic analyses, and X‐ray structural studies. In addition, two ternary complexes of the enzyme with bound S‐adenosyl‐L ‐homocysteine and either dTDP‐3‐amino‐2,3,6‐trideoxy‐4‐keto‐D ‐glucose or dTDP‐3‐amino‐2,3,6‐trideoxy‐D ‐galactose were determined to 1.5 or 1.6 Å resolution, respectively. Taken together, these investigations highlight the important role of His 225 in methyl transfer. In addition, the structural data suggest that the methylation reaction occurs via retention of configuration about the C‐3′ carbon of the sugar.  相似文献   

14.
Bacillus subtilis aminopeptidase hydrolyzed amino acid amides with a specificity similar to that determined using amino acyl-β-naphthylamides, but at much greater catalytic rates. Neutral and basic amino acid amides were the best substrates. A series of Leu and Lys NH2-terminal dipeptides hydrolyzed by Co2+-activated aminopeptidase showed that the kcatKm ratios for the Lys substrates were fourfold greater than the corresponding Leu substrates and that catalytic differences reflected the identity of COOH terminal residues. Greatest catalytic rates were obtained when aromatic residues were in the COOH terminal position of the substrate (Trp, Tyr, Phe); but, significant hydrolysis was achieved when aliphatic residues were COOH-terminal in the dipeptide. The Co2+-activated enzyme would not hydrolyze peptide bonds composed of the imide nitrogen of Pro, thus, bradykinin was not a substrate. However, the Co2+-activated enzyme removed sequentially the first four residues from eledoisin-related peptide and the A chain of bovine insulin.  相似文献   

15.
Proper protein anchoring is key to the biogenesis of prokaryotic cell surfaces, dynamic, resilient structures that play crucial roles in various cell processes. A novel surface protein anchoring mechanism in Haloferax volcanii depends upon the peptidase archaeosortase A (ArtA) processing C‐termini of substrates containing C‐terminal tripartite structures and anchoring mature substrates to the cell membrane via intercalation of lipid‐modified C‐terminal amino acid residues. While this membrane protein lacks clear homology to soluble sortase transpeptidases of Gram‐positive bacteria, which also process C‐termini of substrates whose C‐terminal tripartite structures resemble those of ArtA substrates, archaeosortases do contain conserved cysteine, arginine and arginine/histidine/asparagine residues, reminiscent of His‐Cys‐Arg residues of sortase catalytic sites. The study presented here shows that ArtAWT‐GFP expressed in trans complements ΔartA growth and motility phenotypes, while alanine substitution mutants, Cys173 (C173A), Arg214 (R214A) or Arg253 (R253A), and the serine substitution mutant for Cys173 (C173S), fail to complement these phenotypes. Consistent with sortase active site replacement mutants, ArtAC173A‐GFP, ArtAC173S‐GFP and ArtAR214A‐GFP cannot process substrates, while replacement of the third residue, ArtAR253A‐GFP retains some processing activity. These findings support the view that similarities between certain aspects of the structures and functions of the sortases and archaeosortases are the result of convergent evolution.  相似文献   

16.
Numerous mutations and covalent modifications of the highly abundant, long‐lived crystallins of the eye lens cause their aggregation leading to progressive opacification of the lens, cataract. The nature and biochemical mechanisms of the aggregation process are poorly understood, as neither amyloid nor native‐state polymers are commonly found in opaque lenses. The βγ‐crystallin fold contains four highly conserved buried tryptophans, which can be oxidized to more hydrophilic products, such as kynurenine, upon UV‐B irradiation. We mimicked this class of oxidative damage using Trp→Glu point mutants of human γD‐crystallin. Such substitutions may represent a model of UV‐induced photodamage—introduction of a charged group into the hydrophobic core generating “denaturation from within.” The effects of Trp→Glu substitutions were highly position dependent. While each was destabilizing, only the two located in the bottom of the double Greek key fold—W42E and W130E—yielded robust aggregation of partially unfolded intermediates at 37°C and pH 7. The αB‐crystallin chaperone suppressed aggregation of W130E, but not W42E, indicating distinct aggregation pathways from damage in the N‐terminal vs C‐terminal domain. The W130E aggregates had loosely fibrillar morphology, yet were nonamyloid, noncovalent, showed little surface hydrophobicity, and formed at least 20°C below the melting temperature of the native β‐sheets. These features are most consistent with domain‐swapped polymerization. Aggregation of partially destabilized crystallins under physiological conditions, as occurs in this class of point mutants, could provide a simple in vitro model system for drug discovery and optimization.  相似文献   

17.
18.
Bian G  Huang H  Zong H  Song L 《Chirality》2012,24(10):825-832
A series of new camphorsulfonylated ligands derived from chiral 2‐amino‐2′‐hydroxy‐1,1′‐binaphthyl (NOBIN) and (+)‐camphorsulfonic acid were synthesized by a short and simple synthetic sequence, and their enantioselective catalytic activities were assessed in the nucleophilic addition reaction of dialkylzinc reagents to aldehydes in the presence of titanium tetraisopropoxide. The most efficient ligand, N‐hydroxycamphorsulfonylated (S)‐NOBIN, gave (S)‐addition products with good yields and up to 87% of ee value. The 1H nuclear magnetic resonance (NMR) and 13C NMR results of the titanium titration experiments on this ligand indicate that the most likely catalytic reactive species involved in this catalytic asymmetric addition is a bimetallic titanium complex. A possible catalytic reaction mechanism is proposed. Chirality 24:825–832, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

19.
The 39‐kDa Escherichia coli enzyme MccB catalyses a remarkable posttranslational modification of the MccA heptapeptide during the biosynthesis of microcin C7 (MccC7), a ‘Trojan horse’ antibiotic. The approximately 260‐residue C‐terminal region of MccB is homologous to ubiquitin‐like protein (UBL) activating enzyme (E1) adenylation domains. Accordingly, MccB‐catalysed C‐terminal MccA‐acyl‐adenylation is reminiscent of the E1‐catalysed activation reaction. However, unlike E1 substrates, which are UBLs with a C‐terminal di‐glycine sequence, MccB's substrate, MccA, is a short peptide with an essential C‐terminal Asn. Furthermore, after an intramolecular rearrangement of MccA‐acyl‐adenylate, MccB catalyses a second, unique reaction, producing a stable phosphoramidate‐linked analogue of acyl‐adenylated aspartic acid. We report six‐crystal structures of MccB in apo, substrate‐, intermediate‐, and inhibitor‐bound forms. Structural and kinetic analyses reveal a novel‐peptide clamping mechanism for MccB binding to heptapeptide substrates and a dynamic‐active site for catalysing dual adenosine triphosphate‐consuming reactions. The results provide insight into how a distinctive member of the E1 superfamily carries out two‐step activation for generating the peptidyl‐antibiotic MccC7.  相似文献   

20.
The Escherichia coli serotype O9a O‐antigen polysaccharide (O‐PS) is a model for glycan biosynthesis and export by the ATP‐binding cassette transporter‐dependent pathway. The polymannose O9a O‐PS is synthesized as a polyprenol‐linked glycan by mannosyltransferase enzymes located at the cytoplasmic membrane. The chain length of the O9a O‐PS is tightly regulated by the WbdD enzyme. WbdD first phosphorylates the terminal non‐reducing mannose of the O‐PS and then methylates the phosphate, stopping polymerization. The 2.2 Å resolution structure of WbdD reveals a bacterial methyltransferase domain joined to a eukaryotic kinase domain. The kinase domain is again fused to an extended C‐terminal coiled‐coil domain reminiscent of eukaryotic DMPK (Myotonic Dystrophy Protein Kinase) family kinases such as Rho‐associated protein kinase (ROCK). WbdD phosphorylates 2‐α‐d ‐mannosyl‐d ‐mannose (2α‐MB), a short mimic of the O9a polymer. Mutagenesis identifies those residues important in catalysis and substrate recognition and the in vivo phenotypes of these mutants are used to dissect the termination reaction. We have determined the structures of co‐complexes of WbdD with two known eukaryotic protein kinase inhibitors. Although these are potent inhibitors in vitro, they do not show any in vivo activity. The structures reveal new insight into O‐PS chain‐length regulation in this important model system.  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