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 15 毫秒
1.
The antimicrobial 40‐amino‐acid‐peptide lucifensin was synthesized by native chemical ligation (NCL) using N‐acylbenzimidazolinone (Nbz) as a linker group. NCL is a method in which a peptide bond between two discreet peptide chains is created. This method has been applied to the synthesis of long peptides and proteins when solid‐phase synthesis is imcompatible. Two models of ligation were developed: [15 + 25] Ala‐Cys and [19 + 21] His‐Cys. The [19 + 21] His‐Cys method gives lower yield because of the lower stability of 18‐peptide‐His‐Nbz‐CONH2 peptide, as suggested by density functional theory calculation. Acetamidomethyl‐deprotection and subsequent oxidation of the ligated linear lucifensin gave a mixture of lucifensin isomers, which differed in the location of their disulfide bridges only. The dominant isomer showed unnatural pairing of cysteines [C1?6], [C3?5], and [C2?4], which limits its ability to form α‐helical structure. The activity of isomeric lucifensin toward Bacillus subtilis, Staphylococcus aureus, and Micrococcus luteus was lower than that of the natural lucifensin. The desired product native lucifensin was prepared from this isomer using a one‐pot reduction with dithiotreitol and subsequent air oxidation in slightly alkaline medium. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
James P. Tam  Qitao Yu 《Biopolymers》1998,46(5):319-327
In biological systems, both proteolysis and aminolysis of amide bonds produce activated intermediates through acyl transfer reactions either inter- or intramolecularly. Protein splicing is an illustrative example that proceeds through a series of catalyzed acyl transfer reactions and culminates at an O- or S-acyl intermediate. This intermediate leads to an uncatalyzed acyl migration to form an amide bond in the spliced product. A ligation method mimicking the uncatalyzed final steps in protein splicing has been developed utilizing the acyl transfer amide-bond feature for the blockwise coupling of unprotected, free peptide segments at methionine (Met). The latent thiol moiety of Met can be exploited using homocysteine at the α-amino terminal position of a free peptide for transthioesterification with another free peptide containing an α-thioester to give an S-acyl intermediate. A subsequent, proximity-driven S- to N-acyl migration of this acyl intermediate spontaneously rearranges to form a homocysteinyl amide bond. S-methylation with excess p-nitrobenezensulfonate yields Met at the ligation site. The methionine ligation is selective and orthogonal, and is usually completed within 4 h when performed at slightly basic pH and under strongly reductive conditions. No side reactions due to acylation were observed with any other α-amines of both peptide segments as seen in the synthesis of parathyroid hormone peptides. Furthermore, cyclic peptide can also be obtained through the same strategy by placing both homocysteine at the amino terminus and the thioester at the carboxyl terminus in an unprotected peptide precursor. These biomimetic ligation strategies hold promise for engineering novel peptides and proteins. © 1998 John Wiley & Sons, Inc. Biopoly 46: 319–327, 1998  相似文献   

3.
Human sex hormone binding globulin (hSHBG) is a serum glycoprotein central to the transport and targeted delivery of sex hormones to steroid‐sensitive tissues. Several molecular mechanisms of action of hSHBG, including the function of its attached glycans remain unknown. Here, we perform a detailed site‐specific characterization of the N‐ and O‐linked glycosylation of serum‐derived hSHBG. MS‐driven glycoproteomics and glycomics combined with exoglycosidase treatment were used in a bottom‐up and top‐down manner to determine glycosylation sites, site‐specific occupancies and monosaccharide compositions, detailed glycan structures, and the higher level arrangement of glycans on intact hSHBG. It was found that serum‐derived hSHBG is N‐glycosylated at Asn351 and Asn367 with average molar occupancies of 85.1 and 95.3%, respectively. Both sites are occupied by the same six sialylated and partly core fucosylated bi‐ and triantennary N‐Glycoforms with lactosamine‐type antennas of the form (±NeuAcα6)Galβ4GlcNAc. N‐Glycoforms of Asn367 were slightly more branched and core fucosylated than Asn351 N‐glycoforms due probably to a more surface‐exposed glycosylation site. The N‐terminal Thr7 was fully occupied by the two O‐linked glycans NeuAcα3Galβ3(NeuAcα6)GalNAc (where NeuAc is N‐acetylneuraminic acid and GalNAc is N‐acetylgalactosamine) and NeuAcα3Galβ3GalNAc in a 1:6 molar ratio. Electrophoretic analysis of intact hSHBG revealed size and charge heterogeneity of the isoforms circulating in blood serum. Interestingly, the size and charge heterogeneity were shown to originate predominantly from differential Asn351 glycan occupancies and N‐glycan sialylation that may modulate the hSHBG activity. To date, this work represents the most detailed structural map of the heterogeneous hSHBG glycosylation, which is a prerequisite for investigating the functional aspects of the hSHBG glycans.  相似文献   

4.
Peptide chemical ligation chemistries, which allow the chemoselective coupling of unprotected peptide fragments, are useful tools for synthesizing native polypeptides or unnatural peptide‐based macromolecules. We show here that the phenylthiocarbonyl group can be easily introduced into peptides on α or ε amino groups using phenylthiochloroformate and standard solid‐phase method. It reacts chemoselectively with cysteinyl peptides to give an alkylthiocarbamate bond. S,N‐shift of the alkylaminocarbonyl group from the Cys side chain to the α‐amino group did not occur. The method was used for linking two peptide chains through their N‐termini, for the synthesis of a cyclic peptide or for the synthesis of di‐ or tetravalent multiple antigenic peptides (MAPs). Thiocarbamate ligation is thus complementary to thioether, thioester or disulfide ligation methods. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

5.
The chemical synthesis of proteins has facilitated functional studies of proteins due to the site‐specific incorporation of post‐translational modifications, labels, and non‐proteinogenic amino acids. Moreover, native chemical ligation provides facile access to proteins by chemical means. However, the application of the native chemical ligation reaction in the synthesis of parallel formats such as protein arrays has been complicated because of the often cumbersome and time‐consuming synthesis of the required peptide thioesters. An Fmoc‐based peptide thioester synthesis with self‐purification on the sulfonamide ‘safety‐catch’ linker widens this bottleneck because HPLC purification can be avoided. The method is based on an on‐resin cyclization–thiolysis reaction sequence. A macrocyclization via the N‐terminus of the full‐length peptide followed by a thiolytic C‐terminal ring opening allows selective detachment of the truncation products and the full‐length peptide. A brief overview of the chemical aspects of this method is provided including the optimization steps and the automation process. Furthermore, the application of the cyclization–thiolysis approach combined with the native chemical ligation reaction in the parallel synthesis of a library of 16 SH3‐domain variants of SHO1 in yeast is described, demonstrating the value of this new technique for the chemical synthesis of protein arrays. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

6.
Sortagging is a versatile method for site‐specific modification of proteins as applied to a variety of in vitro reactions. Here, we explore possibilities of adapting the sortase method for use in living cells. For intracellular sortagging, we employ the Ca2+‐independent sortase A transpeptidase (SrtA) from Streptococcus pyogenes. Substrate proteins were equipped with the C‐terminal sortase‐recognition motif (LPXTG); we used proteins with an N‐terminal (oligo)glycine as nucleophiles. We show that sortase‐dependent protein ligation can be achieved in Saccharomyces cerevisiae and in mammalian HEK293T cells, both in the cytosol and in the lumen of the endoplasmic reticulum (ER). ER luminal sortagging enables secretion of the reaction products, among which circular polypeptides. Protein ligation of substrate and nucleophile occurs within 30 min of translation. The versatility of the method is shown by protein ligation of multiple substrates with green fluorescent protein‐based nucleophiles in different intracellular compartments.  相似文献   

7.
Native chemical ligation is widely used for the convergent synthesis of proteins. The peptide thioesters required for this process can be challenging to produce, particularly when using Fmoc‐based solid‐phase peptide synthesis. We have previously reported a route to peptide thioesters, following Fmoc solid‐phase peptide synthesis, via an NS acyl shift that is initiated by the presence of a C‐terminal cysteine residue, under mildly acidic conditions. Under typical reaction conditions, we occasionally observed significant thioester hydrolysis as a consequence of long reaction times (~48 h) and sought to accelerate the reaction. Here, we present a faster route to peptide thioesters, by replacing the C‐terminal cysteine residue with selenocysteine and initiating thioester formation via an NSe acyl shift. This modification allows thioester formation to take place at lower temperatures and on shorter time scales. We also demonstrate how application of this strategy also accelerates peptide cyclization, when a linear precursor is furnished with an N‐terminal cysteine and C‐terminal selenocysteine. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
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   

9.
Native chemical ligation of unprotected peptides in organic solvents has been previously reported as a fast, efficient, and suitable method for coupling of hydrophobic peptides. However, it has not been determined whether the reaction can be carried out without possible side reactions or racemization. Here, we present a study on the chemoselectivity of this method by model reactions designed to test the reactivity of Arg and Lys side chains as well as that of α‐amino groups. A possible racemization of the C‐terminal amino acid of the N‐terminal peptide was also investigated. The results show that ligation in organic solvents can be conducted chemoselectively without side reactions with other nucleophilic groups. Furthermore, no racemization of the C‐terminal amino acid was observed if both educts were added simultaneously. Thus, native chemical ligation can be performed either in aqueous buffer systems or in organic solvents paving the way for the synthesis of larger hydrophobic peptides and/or membrane proteins. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
The membrane‐bound rat GH‐R and an alternatively spliced isoform, the soluble rat GH‐BP, are comprised of identical N‐terminal GH‐binding domains; however, their C‐terminal sequences differ. Immunological reagents are needed to distinguish between the two isoforms in order to understand their respective roles in mediating the actions of GH. Accordingly, a tetravalent MAP dendrimer with four identical branches of a C‐terminal peptide sequence of the rat GH‐BP (GH‐BP263–279) was synthesized and used as an immunogen in rabbits. Solid‐phase peptide synthesis of four GH‐BP263–279 segments onto a tetravalent Lys2‐Lys‐β‐Ala‐OH core peptide was carried out using Fmoc chemistry. The mass of the RP‐HPLC‐purified synthetic product, 8398 Da, determined by ESI‐MS, was identical to expected mass. Three anti‐rat GH‐BP263–279 MAP antisera, BETO‐8039, BETO‐8040, and BETO‐8041, at dilutions of 10?3, recognized both the rat GH‐BP263–279 MAP and recombinant mouse GH‐BP with ED50s within a range of 5–10 fmol, but did not cross‐react with BSA in dot blot analyses. BETO‐8041 antisera (10?3 dilution) recognized GH‐BPs of rat serum and liver having Mrs ranging from 35 to 130 kDa, but did not recognize full‐length rat GH‐Rs. The antisera also detected recombinant mouse GH‐BPs. In summary, the tetravalent rat GH‐BP263–279 MAP dendrimer served as an effective immunogenic antigen in eliciting high titer antisera specific for the C‐termini of both rat and mouse GH‐BPs. The antisera will facilitate studies aimed at improving our understanding of the biology of GH‐BPs. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
Native chemical ligation (NCL) approaches have been applied extensively to soluble proteins. Fewer successes have been achieved with membrane peptides. In this report, the synthesis and semisynthesis by NCL of peptides corresponding to 1.7 transmembrane domains of the α-factor receptor from Saccharomyces cerevisiae is described. Synthesis was achieved when the ligation point was approximately in the middle of the loop joining the two transmembrane regions. In contrast, little to no ligation was observed when the ligation point was at the putative membrane interface of the sixth transmembrane domain (TM6) and the third extracellular loop (EL3). Ligations of a chemically synthesized 22-residue thioester with a synthetic 29-residue N-Cys peptide and a biosynthetic 73-residue N-Cys peptide were successfully achieved in both trifluoroethanol/guanidinium hydrochloride (TFE/GnHCl) and sodium dodecyl sulfate (SDS) media when mercaptoethanesulfonic acid (MESNA) was used as a catalyst. The resulting 51-residue and 95-residue ligation products were purified by reversed phase HPLC and recovered on a mg scale. Both peptides were >95% pure as determined by HPLC and had the expected molecular weight as judged by mass spectrometry. Segmental labeling of the 95-residue fragment, in which the N-Cys portion was [15N] labeled, resulted in a peptide that gave an NMR spectrum which was comparable to that of the unligated 73-residue peptide alone. R B Merrifield personified the finest qualities of a human being. He was an outstanding individual who influenced the way research is conducted by tens of thousands of scientists. At the same time he was a warm, humble, sincere man who was extremely kind and generous. I (FN) personally saw his generosity during a seminar he invited me to give at Rockefeller University. He was already a Nobel laureate but he treated me as a colleague and the encouragement he offered concerning my research program was very important for my future in academia. It is an honor to be among the participants in a volume honoring his contributions to peptide science.  相似文献   

12.
The N‐terminal 1–34 segments of both parathyroid hormone (PTH) and parathyroid hormone‐related protein (PTHrP) bind and activate the same membrane receptor in spite of major differences in their amino acid sequence. The hypothesis was made that they share the same bioactive conformation when bound to the receptor. A common structural motif in all bioactive fragments of the hormone in water/trifluoroethanol mixtures or in aqueous solution containing detergent micelles is the presence of two helical segments at the N‐ and C‐termini of the sequence. In order to stabilize the helical structures, we have recently synthesized and studied the PTHrP(1–34) analog [(Lys13–As p17, Lys26–As p30)]PTHrP(1–34)NH2, which contains lactam‐constrained Lys‐Asp side chains at positions i, i+4. This very potent agonist exhibits enhanced helix stability with respect to the corresponding linear peptide and also two flexible sites at positions 12 and 19 in 1:1 trifluoroethanol/water. These structural elements have been suggested to play a critical role in bioactivity. In the present work we have extended our conformational studies on the bicyclic lactam‐constrained analog to aqueous solution. By CD, 2D‐NMR and structure calculations we have shown that in water two helical segments are present in the region of the lactam bridges (13–18, and 26–31) with high flexibility around Gly12 and Arg19. Thus, the essential structural features observed in the aqueous‐organic medium are maintained in water even if, in this solvent, the overall structure is more flexible. Our findings confirm the stabilizing effect of side‐chain lactam constraints on the α‐helical structure. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
O‐Acyl isopeptides, in which the N‐acyl linkage on the hydroxyamino acid residue (e.g. Ser and Thr) is replaced by an O‐acyl linkage, generally suppress unfavorable aggregation properties derived from the corresponding parent peptides. Here, we report the synthesis of an O‐acyl isopeptide of 34‐mer pyroGlu‐ADan (2), a component of amyloid deposits in hereditary familial Danish dementia, by using native chemical ligation. Native chemical ligation of pyroGlu1‐ADan(1‐21)‐SCH2CH2SO3?Na+ (3) and Cys22O‐acyl isopeptide (4), in which the amino group of the Ser29 residue at the isopeptide moiety was protected by an allyloxycarbonyl group, proceeded well in an aqueous solvent to yield a ligated O‐acyl isopeptide (5). Subsequent disulfide bond formation and deprotection of the allyloxycarbonyl group followed by HPLC purification gave 2 with a reasonable overall yield. 2 was converted to the parent peptide 1 via an O‐to‐N acyl migration reaction. The sequential method, namely (i) native chemical ligation of the O‐acyl isopeptide, (ii) HPLC purification as the O‐acyl isopeptide form, and (iii) O‐to‐N acyl migration into the desired polypeptide, would be helpful to solve problems with HPLC purification of hydrophobic polypeptides in the process of chemical protein synthesis. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
Peptides mimicking antigenic epitopes targeted by antibodies can be powerful tools to be used as antigen surrogates for the specific diagnosis and treatment of autoimmune diseases. Obtaining structural insights about the nature of peptide–antibody interaction in complex mixtures such as sera is a critical goal. In multiple sclerosis (MS), we previously demonstrated that the N‐linked β‐d ‐glucopyranosyl moieties (N‐Glc) containing epitopes in nontypeable Haemophilus influenzae adhesin C‐terminal portion HMW1(1205–1526) were essential for high‐affinity antibody binding in a subpopulation of MS patients. With the aim of developing peptide probes and assessing their binding properties to antibodies from sera of representative patients, we performed the systematic analysis of synthetic peptides based on HMW1(1347–1354) fragment bearing one or two N‐Glc respectively on Asn‐1349 and/or Asn‐1352. The N‐glucosylated nonapeptides efficiently bind to IgG antibodies, displaying IC50 in the range 10?8–10?10 M by competitive indirect enzyme‐linked immunosorbent assay (ELISA) in three representative MS patient sera. We selected the di‐N‐glucosylated adhesin peptide Ac‐KAN (Glc)VTLN (Glc)TT‐NH2 as the shortest sequence able to inhibit high‐avidity interaction with N‐Glc targeting IgM antibodies. Nuclear magnetic resonance (NMR)‐ and circular dichroism (CD)‐based characterization showed that the binding properties of these antigens could not be ascribed to structural differences induced by the presence of up to two N‐glucosyl moieties. Therefore, the antibody binding is not easily correlated to the position of the sugar or to a determined conformation in water.  相似文献   

15.
The terminal homologation by CH2 insertion into the peptides mentioned in the title is described. This involves replacement of the N‐terminal amino acid residue by a β2‐ and of the C‐terminal amino acid residue by a β3‐homo‐amino acid moiety (β2hXaa and β3hXaa, resp.; Fig. 1). In this way, the structure of the peptide chain from the N‐terminal to the C‐terminal stereogenic center is identical, and the modified peptide is protected against cleavage by exopeptidases (Figs. 2 and 3). Neurotensin (NT; 1 ) and its C‐terminal fragment NT(8–13) are ligands of the G‐protein‐coupled receptors (GPCR) NT1, NT2, NT3, and NT analogs are promising tools to be used in cancer diagnostics and therapy. The affinities of homologated NT analogs, 2b – 2e , for NT1 and NT2 receptors were determined by using cell homogenates and tumor tissues (Table 1); in the latter experiments, the affinities for the NT1 receptor are more or less the same as those of NT (0.5–1.3 vs. 0.6 nM ). At the same time, one of the homologated NT analogs, 2c , survives in human plasma for 7 days at 37° (Fig. 6). An NMR analysis of NT(8–13) (Tables 2 and 4, and Fig. 8) reveals that this N‐terminal NT fragment folds to a turn in CD3OH. – In the case of the human analgesic opiorphin ( 3a ), a pentapeptide, and of the HIV‐derived B27‐KK10 ( 4a ), a decapeptide, terminal homologation (→ 3b and 4b , resp.) led to a 7‐ and 70‐fold half‐life increase in plasma (Fig. 9). With N‐terminally homologated NPY, 5c , we were not able to determine serum stability; the peptide consisting of 36 amino acid residues is subject to cleavage by endopetidases. Three of the homologated compounds, 2b, 2c , and 5c , were shown to be agonists (Fig. 7 and 11). A comparison of terminal homologation with other stability‐increasing terminal modifications of peptides is performed (Fig. 5), and possible applications of the neurotensin analogs, described herein, are discussed.  相似文献   

16.
Solid‐state NMR‐based structure determination of membrane proteins and large protein complexes faces the challenge of limited spectral resolution when the proteins are uniformly 13C‐labeled. A strategy to meet this challenge is chemical ligation combined with site‐specific or segmental labeling. While chemical ligation has been adopted in NMR studies of water‐soluble proteins, it has not been demonstrated for membrane proteins. Here we show chemical ligation of the influenza M2 protein, which contains a transmembrane (TM) domain and two extra‐membrane domains. The cytoplasmic domain, which contains an amphipathic helix (AH) and a cytoplasmic tail, is important for regulating virus assembly, virus budding, and the proton channel activity. A recent study of uniformly 13C‐labeled full‐length M2 by spectral simulation suggested that the cytoplasmic tail is unstructured. To further test this hypothesis, we conducted native chemical ligation of the TM segment and part of the cytoplasmic domain. Solid‐phase peptide synthesis of the two segments allowed several residues to be labeled in each segment. The post‐AH cytoplasmic residues exhibit random‐coil chemical shifts, low bond order parameters, and a surface‐bound location, thus indicating that this domain is a dynamic random coil on the membrane surface. Interestingly, the protein spectra are similar between a model membrane and a virus‐mimetic membrane, indicating that the structure and dynamics of the post‐AH segment is insensitive to the lipid composition. This chemical ligation approach is generally applicable to medium‐sized membrane proteins to provide site‐specific structural constraints, which complement the information obtained from uniformly 13C, 15N‐labeled proteins.  相似文献   

17.
Backbone dynamics and conformational properties of drug peptide salmon calcitonin have been studied in aqueous solution using nuclear magnetic resonance (NMR). Although salmon calcitonin (sCT) is largely unfolded in solution (as has been reported in several circular dichroism studies), the secondary Hα chemical shifts and three bond HN–Hα coupling constants indicated that most of the residues of the peptide are populating the α‐helical region of the Ramachandran (?, ψ) map. Further, the peptide in solution has been found to exhibit multiple conformational states exchanging slowly on the NMR timescale (102–103 s?1), inferred by the multiple chemical shift assignments in the region Leu4–Leu12 and around Pro23 (for residues Gln20–Tyr22 and Arg24). Possibly, these slowly exchanging multiple conformational states might inhibit symmetric self‐association of the peptide and, in part, may account for its reduced aggregation propensity compared with human calcitonin (which lacks this property). The 15N NMR‐relaxation data revealed (i) the presence of slow (microsecond‐to‐millisecond) timescale dynamics in the N‐terminal region (Cys1–Ser5) and core residues His17 and Asn26 and (ii) the presence of high frequency (nanosecond‐to‐picosecond) motions in the C‐terminal arm. Put together, the various results suggested that (i) the flexible C‐terminal of sCT (from Thr25–Thr31) is involved in identification of specific target receptors, (ii) whereas the N‐terminal of sCT (from Cys1–Gln20) in solution – exhibiting significant amount of conformational plasticity and strong bias towards biologically active α‐helical structure – facilitates favorable conformational adaptations while interacting with the intermembrane domains of these target receptors. Thus, we believe that the structural and dynamics features of sCT presented here will be useful guiding attributes for the rational design of biologically active sCT analogs. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

18.
In Escherichia coli the Min system prevents Z ring assembly at cell poles by topologically regulating the division inhibitor MinC. The MinC protein has two domains of equal size and both domains can target FtsZ and block cell division in the proper context. Recently, we have shown that, along with MinD, the C‐terminal domain of MinC (MinCC) competes with FtsA, and to a lesser extent with ZipA, for interaction with the C‐terminal tail of FtsZ to block division. Here we explored the interaction between the N‐terminal domain of MinC (MinCN) and FtsZ. A search for mutations in ftsZ that confer resistance to MinCN identified an α‐helix at the interface of FtsZ subunits as being critical for the activity of MinCN. Focusing on one such mutant FtsZ–N280D, we showed that it greatly reduced the FtsZ–MinC interaction and was resistant to MinCN both in vivo and in vitro. With these results, an updated model for the action of MinC on FtsZ is proposed: MinC interacts with FtsZ to disrupt two interactions, FtsZ–FtsA/ZipA and FtsZ–FtsZ, both of which are essential for Z ring formation.  相似文献   

19.
In contrast to the general tendency of hydrophobicity-toxicity relationship of amyloid ?? peptide, we have previously found that the replacement of Asn27 of amyloid ??(25?C35) peptide with Ala yielded a more hydrophobic but less toxic analog and that of Met35 gave a less hydrophobic but more toxic one. To reveal the unique role of these two residues in the neurotoxicity of amyloid ??(1?C42) peptide, the major peptide constituent of amyloid plaques in human brain, we synthesized two analogs N27A and M35A in which Asn27 and Met35 of amyloid ??(1?C42) peptide was replaced with Ala, respectively. The former showed much weaker toxicity than the native peptide, while the latter showed almost an equivalent toxicity, indicating that the side chain amide group of Asn27 has an essential role for the toxicity of amyloid ?? peptides.  相似文献   

20.
Monellin, a sweet protein, consists of two noncovalently associated polypeptide chains: an A chain of 44 amino acid residues and a B chain of 50 residues. Microbial transglutaminase (MTGase) was used for ligation of the monellin subunits without any protecting groups, and without activation of the Cα‐carboxyl group at the C‐terminus. Since a peptide fragment LLQG is a good substrate for MTGase to form an amide bond between the γ‐amide group of the Gln residue and the ε‐amino group of Lys, a monellin B chain analogue in which LLQG was elongated at the C‐terminus (B‐LLQG) was synthesized by solid‐phase synthesis. The monellin A chain analogue in which KGK was elongated at the N‐terminus (KGK‐A) was synthesized by the same method as that of the B chain analogue. The KGK‐A chain and the B‐LLQG chain were coupled by MTGase to give single‐chain analogue of monellin. The single‐chain analogue of monellin was characterized by analytical reverse phase high performance liquid chromatography, electrospray ionization, and amino acid analyses. All analyses gave satisfactory results. The single‐chain analogue of monellin was more heat stable than natural monellin. © 1999 John Wiley & Sons, Inc. Biopoly 50: 193–200, 1999  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