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.
In this study are reported the syntheses of three bis(diarylhydrazonecarbonyl)methylene derivatives [{ArPhCNNH C(O)}2CH2] [Ar = 2 C5H4N (5), C6H5 (6), and 2‐C4H3S (7)], obtained by condensation of corresponding hydrazones with carbon suboxide, C3O2. The solid‐state self‐assembly of these carbonyl derivatives, giving rise to polymeric and dimeric networks, is described. In the formation of these structural features, in addition to N—H· · ·OC intermolecular hydrogen bonds, stabilizing intramolecular NH· · · π (systems) and intermolecular CO· · ·π (systems) interactions also seem to play an important role. Solution 1H‐nmr data of compounds 5–7 indicate that the polymeric and dimeric structures are not maintained in solution and show the occurrence of keto‐enolic equilibria. © 1999 John Wiley & Sons, Inc. Biopoly 49: 541–549, 1999  相似文献   

2.
Elucidating protein function from its structure is central to the understanding of cellular mechanisms. This involves deciphering the dependence of local structural motifs on sequence. These structural motifs may be stabilized by direct or water‐mediated hydrogen bonding among the constituent residues. π‐Turns, defined by interactions between (i) and (i + 5) positions, are large enough to contain a central space that can embed a water molecule (or a protein moiety) to form a stable structure. This work is an analysis of such embedded π‐turns using a nonredundant dataset of protein structures. A total of 2965 embedded π‐turns have been identified, as also 281 embedded Schellman motif, a type of π‐turn which occurs at the C‐termini of α‐helices. Embedded π‐turns and Schellman motifs have been classified on the basis of the protein atoms of the terminal turn residues that are linked by the embedded moiety, conformation, residue composition, and compared with the turns that have terminal residues connected by direct hydrogen bonds. Geometrically, the turns have been fitted to a circle and the position of the linker relative to its center analyzed. The hydroxyl group of Ser and Thr, located at (i + 3) position, is the most prominent linker for the side‐chain mediated π‐turns. Consideration of residue conservation among homologous sequences indicates the terminal and the linker positions to be the most conserved. The embedded π‐turn as a binding site (for the linker) is discussed in the context of “nest,” a concave depression that is formed in protein structures with adjacent residues having enantiomeric main‐chain conformations. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 441–453, 2014.  相似文献   

3.
Young Kee Kang    Byung Jin Byun 《Biopolymers》2012,97(12):1018-1025
The γ‐peptide β‐turn structures have been designed computationally by the combination of chirospecific γ 2 , 3 ‐residues of 2‐(aminomethyl)cyclohexanecarboxylic acid (γAmc6) with a cyclohexyl constraint on the Cα?Cβ bond using density functional methods in water. The chirospecific γAmc6 dipeptide with the (2S,3S)‐(2R,3R) configurations forms a stable turn structure in water, resembling a type II′ turn of α‐peptides, which can be used as a β‐turn motif in β‐hairpins of Ala‐based α‐peptides. The γAmc6 dipeptide with homochiral (2S,3S)‐(2S,3S) configurations but different cyclohexyl puckerings shows the capability to be incorporated into one of two β‐turn motifs of gramicidin S. The overall structure of this gramicidin S analogue is quite similar to the native gramicidin S with the same patterns and geometries of hydrogen bonds. Our calculated results and the recently observed results may imply the wider applicability of chirospecific γ‐peptides with a cyclohexyl constraint on the backbone to form various peptide foldamers. © 2012 Wiley Periodicals, Inc. Biopolymers 97:1018–1025, 2012.  相似文献   

4.
To investigate the structural role played by isostructural unbranched alkyl‐chains on the conformational ensemble and stability of β‐turn structures, the conformational properties of a designed model peptide: Plm‐Pro‐Gly‐Pda ( 1 , Plm: H3C—(CH2)14—CONH—; Pda: —CONH— (CH2)14—CH3) have been examined and compared with the parent peptide: Boc‐Pro‐Gly‐NHMe ( 2 , Boc: tert‐butoxycarbonyl; NHMe: N‐methylamide). The characteristic 13C NMR chemical‐shifts of the Pro Cβ and Cγ resonances ascertained the incidence of an all‐trans peptide‐bond in low polarity deuterochloroform solution. Using FTIR and 1H NMR spectroscopy, we establish that apolar alkyl‐chains flanking a β‐turn promoting Pro‐Gly sequence impart definite incremental stability to the well‐defined hydrogen‐bonded structure. The assessment of 1H NMR derived thermodynamic parameters of the hydrogen‐bonded amide‐NHs via variable temperature indicate that much weaker hydrophobic interactions do contribute to the stability of folded reverse turn structures. The far‐UV CD spectral patterns of 1 and 2 in 2,2,2‐trifluoroethanol are consistent with Pro‐Gly specific type II β‐turn structure, concomitantly substantiate that the flanking alkyl‐chains induce substantial bias in enhanced β‐turn populations. In view of structural as well as functional importance of the Pro‐Gly mediated secondary structures, besides biochemical and biological significance of proteins lipidation via myristoylation or palmytoilation, we highlight potential convenience of the unbranched Plm and Pda moieities not only as main‐chain N‐ and C‐terminal protecting groups but also to mimic and stabilize specific isolated secondary and supersecondary structural components frequently observed in proteins and polypeptides. © 2013 Wiley Periodicals, Inc. Biopolymers 99: 419–426, 2013.  相似文献   

5.
Our goal was to gain a better understanding of how protein stability can be increased by improving β‐turns. We studied 22 β‐turns in nine proteins with 66–370 residues by replacing other residues with proline and glycine and measuring the stability. These two residues are statistically preferred in some β‐turn positions. We studied: Cold shock protein B (CspB), Histidine‐containing phosphocarrier protein, Ubiquitin, Ribonucleases Sa2, Sa3, T1, and HI, Tryptophan synthetase α‐subunit, and Maltose binding protein. Of the 15 single proline mutations, 11 increased stability (Average = 0.8 ± 0.3; Range = 0.3–1.5 kcal/mol), and the stabilizing effect of double proline mutants was additive. On the basis of this and our previous work, we conclude that proteins can generally be stabilized by replacing nonproline residues with proline residues at the i + 1 position of Type I and II β‐turns and at the i position in Type II β‐turns. Other turn positions can sometimes be used if the φ angle is near ?60° for the residue replaced. It is important that the side chain of the residue replaced is less than 50% buried. Identical substitutions in β‐turns in related proteins give similar results. Proline substitutions increase stability mainly by decreasing the entropy of the denatured state. In contrast, the large, diverse group of proteins considered here had almost no residues in β‐turns that could be replaced by Gly to increase protein stability. Improving β‐turns by substituting Pro residues is a generally useful way of increasing protein stability. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

6.
The conformation of oligomers of β‐amino acids of the general type Ac‐[β‐Xaa]n‐NHMe (β‐Xaa = β‐Ala, β‐Aib, and β‐Abu; n = 1–4) was systematically examined at different levels of ab initio molecular orbital theory (HF/6‐31G*, HF/3‐21G). The solvent influence was considered employing two quantum‐mechanical self‐consistent reaction field models. The results show a wide variety of possibilities for the formation of characteristic elements of secondary structure in β‐peptides. Most of them can be derived from the monomer units of blocked β‐peptides with n = 1. The stability and geometries of the β‐peptide structures are considerably influenced by the side‐chain positions, by the configurations at the Cα‐ and Cβ‐atoms of the β‐amino acid constituents, and especially by environmental effects. Structure peculiarities of β‐peptides, in particular those of various helix alternatives, are discussed in relation to typical elements of secondary structure in α‐peptides. © 1999 John Wiley & Sons, Inc. Biopoly 50: 167–184, 1999  相似文献   

7.
The increasing interest in click chemistry and its use to stabilize turn structures led us to compare the propensity for β‐turn stabilization of different analogs designed as mimics of the β‐turn structure found in tendamistat. The β‐turn conformation of linear β‐amino acid‐containing peptides and triazole‐cyclized analogs were compared to ‘conventional’ lactam‐ and disulfide‐bridged hexapeptide analogs. Their 3D structures and their propensity to fold in β‐turns in solution, and for those not structured in solution in the presence of α‐amylase, were analyzed by NMR spectroscopy and by restrained molecular dynamics with energy minimization. The linear tetrapeptide Ac‐Ser‐Trp‐Arg‐Tyr‐NH2 and both the amide bond‐cyclized, c[Pro‐Ser‐Trp‐Arg‐Tyr‐D ‐Ala] and the disulfide‐bridged, Ac‐c[Cys‐Ser‐Trp‐Arg‐Tyr‐Cys]‐NH2 hexapeptides adopt dominantly in solution a β‐turn conformation closely related to the one observed in tendamistat. On the contrary, the β‐amino acid‐containing peptides such as Ac‐(R)‐β3‐hSer‐(S)‐Trp‐(S)‐β3‐hArg‐(S)‐β3‐hTyr‐NH2, and the triazole cyclic peptide, c[Lys‐Ser‐Trp‐Arg‐Tyr‐βtA]‐NH2, both specifically designed to mimic this β‐turn, do not adopt stable structures in solution and do not show any characteristics of β‐turn conformation. However, these unstructured peptides specifically interact in the active site of α‐amylase, as shown by TrNOESY and saturation transfer difference NMR experiments performed in the presence of the enzyme, and are displaced by acarbose, a specific α‐amylase inhibitor. Thus, in contrast to amide‐cyclized or disulfide‐bridged hexapeptides, β‐amino acid‐containing peptides and click‐cyclized peptides may not be regarded as β‐turn stabilizers, but can be considered as potential β‐turn inducers. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
Hyun Joo  Jerry Tsai 《Proteins》2014,82(9):2128-2140
To understand the relationship between protein sequence and structure, this work extends the knob‐socket model in an investigation of β‐sheet packing. Over a comprehensive set of β‐sheet folds, the contacts between residues were used to identify packing cliques: sets of residues that all contact each other. These packing cliques were then classified based on size and contact order. From this analysis, the two types of four‐residue packing cliques necessary to describe β‐sheet packing were characterized. Both occur between two adjacent hydrogen bonded β‐strands. First, defining the secondary structure packing within β‐sheets, the combined socket or XY:HG pocket consists of four residues i, i+2 on one strand and j, j+2 on the other. Second, characterizing the tertiary packing between β‐sheets, the knob‐socket XY:H+B consists of a three‐residue XY:H socket (i, i+2 on one strand and j on the other) packed against a knob B residue (residue k distant in sequence). Depending on the packing depth of the knob B residue, two types of knob‐sockets are found: side‐chain and main‐chain sockets. The amino acid composition of the pockets and knob‐sockets reveal the sequence specificity of β‐sheet packing. For β‐sheet formation, the XY:HG pocket clearly shows sequence specificity of amino acids. For tertiary packing, the XY:H+B side‐chain and main‐chain sockets exhibit distinct amino acid preferences at each position. These relationships define an amino acid code for β‐sheet structure and provide an intuitive topological mapping of β‐sheet packing. Proteins 2014; 82:2128–2140. © 2014 Wiley Periodicals, Inc.  相似文献   

9.
Both Type I' and Type II' β‐turns have the same sense of the β‐turn twist that is compatible with the β‐sheet twist. They occur predominantly in two residue β‐hairpins, but the occurrence of Type I' β‐turns is two times higher than Type II' β‐turns. This suggests that Type I' β‐turns may be more stable than Type II' β‐turns, and Type I' β‐turn sequence and structure can be more favorable for protein folding than Type II' β‐turns. Here, we redesigned the native Type II' β‐turn in GFP to Type I' β‐turn, and investigated its effect on protein folding and stability. The Type I' β‐turns were designed based on the statistical analysis of residues in natural Type I' β‐turns. The substitution of the native “GD” sequence of i+1 and i+2 residues with Type I' preferred “(N/D)G” sequence motif increased the folding rate by 50% and slightly improved the thermodynamic stability. Despite the enhancement of in vitro refolding kinetics and stability of the redesigned mutants, they showed poor soluble expression level compared to wild type. To overcome this problem, i and i + 3 residues of the designed Type I' β‐turn were further engineered. The mutation of Thr to Lys at i + 3 could restore the in vivo soluble expression of the Type I' mutant. This study indicates that Type II' β‐turns in natural β‐hairpins can be further optimized by converting the sequence to Type I'. Proteins 2014; 82:2812–2822. © 2014 Wiley Periodicals, Inc.  相似文献   

10.
The synthesis of new dermorphin analogues is described. The (R)‐alanine or phenylalanine residues of natural dermorphin were substituted by the corresponding α‐methyl‐β‐azidoalanine or α‐benzyl‐β‐azido(1‐piperidinyl)alanine residues. The potency and selectivity of the new analogues were evaluated by a competitive receptor binding assay in rat brain using [3H]DAMGO (a μ ligand) and [3H]DELT (a δ ligand). The most active analogue in this series, Tyr‐(R)‐Ala‐(R)‐α‐benzyl‐β‐azidoAla‐Gly‐Tyr‐Pro‐Ser‐NH2 and its epimer were analysed by 1H and 13C NMR spectroscopy and restrained molecular dynamics simulations. The dominant conformation of the investigated peptides depended on the absolute configuration around Cα in the α‐benzyl‐β‐azidoAla residue in position 3. The (R) configuration led to the formation of a type I β‐turn, whilst switching to the (S) configuration gave rise to an inverse β‐turn of type I′, followed by the formation of a very short β‐sheet. The selectivity of Tyr‐(R)‐Ala‐(R) and (S)‐α‐benzyl‐β‐azidoAla‐Gly‐Tyr‐Pro‐Ser‐NH2 was shown to be very similar; nevertheless, the two analogues exhibited different conformational preferences. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
Young Kee Kang  In Kee Yoo 《Biopolymers》2014,101(11):1077-1087
Conformational preferences of 9‐ and 14‐helix foldamers have been studied for γ‐dipeptides of 2‐aminocyclohexylacetic acid (γAc6a) residues such as Ac‐(γAc6a)2‐NHMe ( 1 ), Ac‐(Cα‐Et‐γAc6a)2‐NHMe ( 2 ), Ac‐(γAc6a)2‐NHBn ( 3 ), and Ac‐(Cα‐Et‐γAc6a)2‐NHBn ( 4 ) at the M06‐2X/cc‐pVTZ//M06‐2X/6‐31 + G(d) level of theory to explore the influence of substituents on their conformational preferences. In the gas phase, the 9‐helix foldamer H9 and 14‐helix foldamer H14‐z are found to be most preferred for dipeptides 2 and 4 , respectively, as for dipeptides 1 and 3 , which indicates no remarkable influence of the Cα‐ethyl substitution on conformational preferences. The benzyl substitution at the C‐terminal end lead H14‐z to be the most preferred conformer for dipeptides 3 and 4 , whereas it is H9 for dipeptides 1 and 2 , which can be ascribed to the favored C? H···π interactions between the cyclohexyl group of the first residue and the C‐terminal benzyl group. There are only marginal changes in backbone structures and the distances and angles of H‐bonds for all local minima by Cα‐ethyl and/or benzyl substitutions. Although vibrational frequencies and intensities of the dipeptide 4 calculated at both M06‐2X/6‐31 + G(d) and M05‐2X/6‐31 + G(d) levels of theory are consistent with observed results in the gas phase, H14‐z is predicted to be most preferred by ΔG only at the former level of theory. Hydration did not bring the significant changes in backbone structures of helix foldamers for both dipeptide 1 and 4 . It is expected that the different substitutions at the C‐terminal end lead to the different helix foldamers, which may increase the resistance of helical structures to proteolysis and provide the more surface to the helical structures suitable for molecular recognition. © 2014 Wiley Periodicals, Inc. Biopolymers 101: 1077–1087, 2014.  相似文献   

12.
Selective agonist of nonopioid β‐endorphin receptor decapeptide immunorphin (SLTCLVKGFY) was labeled with tritium (the specific activity of 24 Ci/mmol). [3H]Immunorphin was found to bind to nonopioid β‐endorphin receptor of mouse peritoneal macrophages (Kd = 2.0 ± 0.1 nM ). The [3H]immunorphin specific binding with macrophages was inhibited by unlabeled β‐endorphin (Ki = 2.9 ± 0.2 nM ) and was not inhibited by unlabeled naloxone, α‐endorphin, γ‐endorphin and [Met5]enkephalin (Ki > 10 µM ). Thirty fragments of β‐endorphin have been synthesized and their ability to inhibit the [3H]immunorphin specific binding to macrophages was studied. Unlabeled fragment 12–19 (TPLVTLFK, the author's name of the peptide octarphin) was found to be the shortest peptide possessing practically the same inhibitory activity as β‐endorphin (Ki = 3.1 ± 0.3 nM ). The peptide octarphin was labeled with tritium (the specific activity of 28 Ci/mmol). [3H]Octarphin was found to bind to macrophages with high affinity (Kd = 2.3 ± 0.2 nM ). The specific binding of [3H]octarphin was inhibited by unlabeled immunorphin and β‐endorphin (Ki = 2.4 ± 0.2 and 2.7 ± 0.2 nM , respectively). Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
The crystal structure of a tripeptide Boc‐Leu‐Val‐Ac12c‐OMe ( 1 ) is determined, which incorporates a bulky 1‐aminocyclododecane‐1‐carboxylic acid (Ac12c) side chain. The peptide adopts a semi‐extended backbone conformation for Leu and Val residues, while the backbone torsion angles of the Cα,α‐dialkylated residue Ac12c are in the helical region of the Ramachandran map. The molecular packing of 1 revealed a unique supramolecular twisted parallel β‐sheet coiling into a helical architecture in crystals, with the bulky hydrophobic Ac12c side chains projecting outward the helical column. This arrangement resembles the packing of peptide helices in crystal structures. Although short oligopeptides often assemble as parallel or anti‐parallel β‐sheet in crystals, twisted or helical β‐sheet formation has been observed in a few examples of dipeptide crystal structures. Peptide 1 presents the first example of a tripeptide showing twisted β‐sheet assembly in crystals. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
The three‐dimensional solution structure of harzianin HC IX, a peptaibol antibiotic isolated from the fungus Trichoderma harzianum, was determined using CD, homonuclear, and heteronuclear two‐dimensional nmr spectroscopy combined with molecular modeling. This 14‐residue peptide, Ac Aib1 Asn2 Leu3 Aib4 Pro5 Ala6 Ile7 Aib8 Pro9 Iva10 Leu11 Aib12 Pro13 Leuol14 (Aib, α‐aminoisobutyric acid; Iva, isovaline; Leuol, leucinol), is a main representative of a short‐sequence peptaibol class characterized by an acetylated N‐terminus, a C‐terminal amino alcohol, and the presence of three Aib‐L ‐Pro motifs at positions 4–5, 8–9, and 12–13, separated by two dipeptide units. In spite of a lower number of residues, compared to the 18/20‐residue peptaibols such as alamethicin, harzianin HC IX exhibits remarkable membrane‐perturbing properties. It interacts with phospholipid bilayers, increasing their permeability and forming voltage‐gated ion channels through a mechanism slightly differing from that proposed for alamethicin. Sequence‐specific 1H‐ and 13C‐nmr assignments and conformational nmr parameters (3JNHCαH coupling constants, quantitative nuclear Overhauser enhancement data, temperature coefficients of amide and carbonyl groups, NH–ND exchange rates) were obtained in methanol solution. Sixty structures were calculated based on 98 interproton distance restraints and 6 Φ dihedral angle restraints, using high temperature restrained molecular dynamics and energy minimization. Thirty‐seven out of the sixty generated structures were consistent with the nmr data and were convergent. The peptide backbone consists in a ribbon of overlapping β‐turns twisted into a continuous spiral from Asn2 to Leuol14 and forming a 26 Å long helix‐like structure. This structure is slightly amphipathic, with the three Aib–Pro motifs aligned on the less hydrophobic face of the spiral where the Asn2 side chain is also present, while the more hydrophobic bulky side chains of leucines, isoleucine, isovaline, and leucinol are located on the concave side. The repetitive (Xaa–Yaa–Aib–Pro) tetrapeptide subunit, making up the peptide sequence, is characterized by four sets of (Φ,Ψ) torsional angles, with the following mean values: Φi = −90°, Ψi = −27°; Φi+1 = −98°, Ψi+1 = −17°; Φi+2 = −49°, Ψi+2 = −50°; Φi+3 = −78°, Ψi+3 = +3°. We term this particular structure, specifically occurring in the case of (Xaa–Yaa–Aib–Pro)n sequences, the (Xaa–Yaa–Aib–Pro)‐β‐bend ribbon spiral. It is stabilized by 4 → 1 intramolecular hydrogen bonds and differs from both the canonical 310‐helix made of a succession of type III β‐turns and from the β‐bend ribbon spiral that has been described in the case of (Aib–Pro)n peptide segments. © 1999 John Wiley & Sons, Inc. Biopoly 50: 71–85, 1999  相似文献   

15.
β‐Amino acids containing hybrid peptides and β‐peptides show great potential as peptidomimetics. In this paper we describe the synthesis and affinity toward the µ‐ and δ‐opioid receptors of β‐peptides, analogues of Leu‐enkephalin, deltorphin I, dermorphin and α,β‐hybrides, analogues of deltorphin I. Substitution of α‐amino acid residues with β3homo‐amino acid residues, in general resulted in decrease of affinity to opioid receptors. However, the incorporation β3h‐D ‐Ala in position 2 or β3hPhe in position 3 of deltorphin I resulted in potent and selective ligand for δ‐opioid receptor. The NMR studies of β‐deltorphin I analogue suggest that conformational motions in the central part of the peptide backbone are partially restricted and some conformational preferences can be expected. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
Aims: To characterize a β‐xylosidase from the thermophilic fungus Thermomyces lanuginosus and to investigate its potential in saccharification of hemicellulosic xylans. Methods and Results: A gene (designated TlXyl43) encoding β‐xylosidase was cloned from T. lanuginosus CAU44 and expressed in Escherichia coli. The gene consists of a 1017‐bp open reading frame without introns. It encodes a mature protein of 338 residues with no predicted signal peptide, belonging to glycoside hydrolase (GH) family 43. Over 60% of the recombinant β‐xylosidase (TlXyl43) was secreted into the culture medium. TlXyl43 was purified 2·6‐fold to homogeneity with an estimated mass of 51·6 kDa by SDS‐PAGE. The purified enzyme exhibited optimal activity at pH 6·5 and 55°C and was stable at 50°C. It was competitively inhibited by xylose with a Ki value of 63 mmol l?1. Conclusions: In this study, a GH family 43 β‐xylosidase gene (TlXyl43) from T. lanuginosus CAU44 was cloned and functionally expressed in E. coli, and over 60% of recombinant protein was secreted into the culture. Significance and Impact of the Study: This is the first report of the cloning and functional expression of a β‐xylosidase gene from Thermomyces species. TlXyl43 holds great potential for variety of industries.  相似文献   

17.
Replica exchange molecular dynamics simulations (300 ns) were used to study the dimerization of amyloid β(1‐40) (Aβ(1‐40)) polypeptide. Configurational entropy calculations revealed that at physiological temperature (310 K, 37°C) dynamic dimers are formed by randomly docked monomers. Free energy of binding of the two chains to each other was ?93.56 ± 6.341 kJ mol?1. Prevalence of random coil conformations was found for both chains with the exceptions of increased β‐sheet content from residues 16‐21 and 29‐32 of chain A and residues 15‐21 and 30‐33 of chain B with β‐turn/β‐bend conformations in both chains from residues 1‐16, 21‐29 of chain A, 1‐16, and 21‐29 of chain B. There is a mixed β‐turn/β‐sheet region from residues 33‐38 of both chains. Analysis of intra‐ and interchain residue distances shows that, although the individual chains are highly flexible, the dimer system stays in a loosely packed antiparallel β‐sheet configuration with contacts between residues 17‐21 of chain A with residues 17‐21 and 31‐36 of chain B as well as residues 31‐36 of chain A with residues 17‐21 and 31‐36 of chain B. Based on dihedral principal component analysis, the antiparallel β‐sheet‐loop‐β‐sheet conformational motif is favored for many low energy sampled conformations. Our results show that Aβ(1‐40) can form dynamic dimers in aqueous solution that have significant conformational flexibility and are stabilized by collapse of the central and C‐terminal hydrophobic cores with the expected β‐sheet‐loop‐β‐sheet conformational motif. Proteins 2017; 85:1024–1045. © 2017 Wiley Periodicals, Inc.  相似文献   

18.
The relationship between the conformation and biological activity of the peptide allosteric modulator of the interleukin‐1 receptor 101.10 (D ‐Arg‐D ‐Tyr‐D ‐Thr‐D ‐Val‐D ‐Glu‐D ‐Leu‐D ‐Ala‐NH2) has been studied using (R)‐ and (S)‐Bgl residues. Twelve Bgl peptides were synthesized using (R)‐ and (S)‐cyclic sulfamidate reagents derived from L ‐ and D ‐aspartic acid in an optimized Fmoc‐compatible protocol for efficient lactam installment onto the supported peptide resin. Examination of these (R)‐ and (S)‐Bgl 101.10 analogs for their potential to inhibit IL‐1β‐induced thymocyte cell proliferation using a novel fluorescence assay revealed that certain analogs exhibited retained and improved potency relative to the parent peptide 101.10. In light of previous reports that Bgl residues may stabilize type II′β‐turn‐like conformations in peptides, CD spectroscopy was performed on selected compounds to identify secondary structure necessary for peptide biological activity. Results indicate that the presence of a fold about the central residues of the parent peptide may be important for activity. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
The synthetic peptide octarphin (TPLVTLFK) corresponding to the sequence 12–19 of β‐endorphin, a selective agonist of nonopioid β‐endorphin receptor, was labeled with tritium to a specific activity of 29 Ci/mmol. [3H]Octarphin was found to bind to high‐affinity naloxone‐insensitive binding sites on membranes isolated from rat adrenal cortex (Kd = 35.7 ± 2.3 nM, Bmax = 41.0 ± 3.6 pmol/mg protein). The binding specificity study revealed that these binding sites were insensitive not only to naloxone but to α‐endorphin, γ‐endorphin, [Met5]enkephalin, and [Leu5]enkephalin as well. At the same time, the [3H]octarphin‐specific binding with adrenal cortex membranes was inhibited by unlabeled β‐endorphin (Ki = 32.9 ± 3.8 nM). Octarphin at concentrations of 10?9–10?6 M was found to inhibit the adenylate cyclase activity in adrenocortical membranes, whereas intranasal injection of octarphin at doses of 5 and 20 µg/rat was found to reduce the secretion of corticosterone from the adrenals to the bloodstream. Thus, octarphin decreases the adrenal cortex functional activity through the high affinity binding to nonopioid receptor of β‐endorphin. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
Guangcai Ma  Wenyou Zhu  Yongjun Liu 《Proteins》2016,84(11):1606-1615
Pectate lyase utilizes the anti‐β‐elimination chemistry to catalyze the cleavage of α‐1,4 glycosidic bond between D‐galacturonate regions during the degradation of plant polysaccharide pectin. We report here detailed mechanistic studies of the Bacillus subtilis pectate lyase (BsPel) using QM/MM calculations. It was found that the residue Arg279 serves as the catalytic base to abstract the α‐proton from C52 atom of substrate Ada2 subsite, forming an unstable carbanion intermediate. The glycosidic bond of this intermediate is scissile to generate the 4,5‐unsaturated digalacturonate product and a negatively charged β‐leaving group. Two active site residues (Lys247 and Arg279) and two Ca2+ ions (Ca2 and Ca3) form hydrogen‐bonding and coordination interactions with C52? COO? of Ada2, respectively, which facilitate the proton abstraction and stabilize the generated carbanion intermediates. Arg284 is not the potential proton donor to saturate the leaving group. Actually, the proton source of leaving group is the solvent water molecule rather than any active site acidic residues. In addition, the calculation results suggest that careful selections of QM‐ and Active‐regions are essential to accurately explore the enzymatic reactions. Proteins 2016; 84:1606–1615. © 2016 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