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.
Abstract

A peptide corresponding to residues 101–141 of the human nucleotide excision repair protein XPA was synthesized with an isoleucine substituted for L138 and its solution structure studied by circular dichroism and homonuclear 1H NMR spectroscopy. The peptide, (XPA-41), contains a C4?type zinc-binding motif, C105-(X)2C108-(X)l7?C126-(X)2 C129, which XPA requires for damaged-DNA binding activity. The proton resonances of XPA-41without zinc (apoXPA-41) were assigned using homonuclear TOCSY, NOESY and DQF-COSY data and show the apo-zinc peptide is a random coil. The peptide was folded with the addition of 1.2 equivalents of ZnCl2 in dilute solution at pH 4.0. Electrospray ionization mass spectroscopy illustrated an increase in the molecular weight of XPA-41 by 65 amu. Circular dichroism spectra of the zinc-folded peptide (zXPA-41) showed the acquisition of elements of secondary structure. Such a conclusion was confirmed with'H NMR data collected at 25°C, pH 6.3. Hα-secondary shifts and NOE patterns indicate that regions V102-C105 and G109-F112 form an anti-parallel β-sheet and residues N128-K137 form a nascent α-helix. Rapid exchange of most amide resonances between S115-C126 prohibited unambiguous assignment of all the proton resonances in this region. However, a 1.19 ppm downfield shift of the Hα resonance of T125 relative to the apo-zinc peptide, together with downfield shifted Hα resonances for the adjacent residues (P124 and L123), suggest a second β-sheet is present in the S115-C126 region. On the basis of structural similarities to GATA-1 (Science 267:438–446), a homology generated structure for zXPA-41 was made, using GATA-1 as the template, which satisfied all the observed NOEs. Using the hybrid homology-NMR based zXPA-41 structure and analogy to GATA-1, models for the role played by the zinc-binding core (101–141) of XPA in DNA damage recognition are proposed.  相似文献   

2.
Abstract: The regulation of adenylate cyclase activity by adrenocorticotropin/α-melanocyte–stimulating hormone (ACTH/MSH)-like peptides was investigated in rat brain slices using a superfusion method. Adenylate cyclase activity was concentration-dependently increased by ACTH-(1–24), α-MSH (EC50 values 16 and 6 nM, respectively), and [Nle4,D-Phe7]α-MSH (EC50 value 1.6 nM), in the presence of forskolin (1 μM, optimal concentration). 1-9-Dideoxy-forskolin did not augment the response of adenylate cyclase to ACTH-(1–24). Various peptide fragments were tested for their ability to enhance [3H]cyclic AMP production. [Nle4,D-Phe7]α-MSH increased [3H]cyclic AMP formation with a maximal effect of 30% and was more potent than ACTH-(1–24), ACTH-(1–16)-NH2, α-MSH, ACTH-(1–13)-NH2, [MetO4]α-MSH, [MetO24,D-Lys8,Phe9]ACTH-(4–9), ACTH-(7–16)-NH2, ACTH-(1–10), and ACTH-(11–24), in order of potency. This structure–activity relationship resembles that found for the previously described peptide-induced display of excessive grooming. ACTH-(1–24) stimulated adenylate cyclase activity in both striatal (maximal effect, ?20%) and septal slices (maximal effect, ?40%), but not in hippocampal or cortical slices. Lesioning of the dopaminergic projections to the striatum did not result in a diminished effect of [Nle4,D-Phe7]α-MSH on [3H]cyclic AMP accumulation, which indicates that the ACTH/MSH receptor–stimulated adenylate cyclase is not located on striatal dopaminergic terminals. ACTH-(1–24) did not affect the dopamine D1 or D2 receptor–mediated modulation of adenylate cyclase activity. Based on the present data, we suggest that the binding of endogenous ACTH or α-MSH to a putative ACTH/MSH receptor in certain brain regions leads to the activation of a signal transduction pathway using cyclic AMP as a second messenger.  相似文献   

3.
Synthetic cyclic octapeptides of general structure cyclo[Glu(γOBzl)-Sar-Gly-(N-R)Gly]2 (R = n-hexyl and cyclohexyl) transport calcium ions selectively across organic phases and phospholipid membranes. We have now used proton nmr spectroscopy (360 MHz) to study the solution conformation(s) of their calcium complexes. When Ca(ClO4)2 was added to solutions of these peptides in CDCl3, nmr spectra of the resulting calcium complexes were characteristic of a single C2-symmetric conformer. From a Karplus-Bystrov analysis of vicinal coupling constants in both the peptide backbone and Glu side chain (treated as an ABCCMX spin system), in conjuction with model-building studies, a structure was proposed in which the calcium ion is bound in an octahedral-type complex by the four (coplanar) carbonyl groups of the (all-trans) Glu-Sar and Gly-(N-R)Gly peptide bonds. Occurrence of preferred rotamers about Glu side chain Cα–Cβ bonds indicated that restricted rotation in peptide side chains arises upon calcium binding.  相似文献   

4.
The solid‐state conformations of two αγ hybrid peptides Boc‐[Aib‐γ4(R)Ile]4‐OMe 1 and Boc‐[Aib‐γ4(R)Ile]5‐OMe 2 are described. Peptides 1 and 2 adopt C12‐helical conformations in crystals. The structure of octapeptide 1 is stabilized by six intramolecular 4 → 1 hydrogen bonds, forming 12 atom C12 motifs. The structure of peptide 2 reveals the formation of eight successive C12 hydrogen‐bonded turns. Average backbone dihedral angles for αγ C12 helices are peptide 1 , Aib; φ (°) = ?57.2 ± 0.8, ψ (°) = ?44.5 ± 4.7; γ4(R)Ile; φ (°) = ?127.3 ± 7.3, θ1 (°) = 58.5 ± 12.1, θ2 (°) = 67.6 ± 10.1, ψ (°) = ?126.2 ± 16.1; peptide 2 , Aib; φ (°) = ?58.8 ± 5.1, ψ (°) = ?40.3 ± 5.5; ψ4(R)Ile; φ (°) = ?123.9 ± 2.7, θ1 (°) = 53.3 θ 4.9, θ 2 (°) = 61.2 ± 1.6, ψ (°) = ?121.8 ± 5.1. The tendency of γ4‐substituted residues to adopt gauche–gauche conformations about the Cα–Cβ and Cβ–Cγ bonds facilitates helical folding. The αγ C12 helix is a backbone expanded analog of α peptide 310 helix. The hydrogen bond parameters for α peptide 310 and α‐helices are compared with those for αγ hybrid C12 helix. Copyright © 2016 European Peptide Society and John Wiley & Sons.  相似文献   

5.
The crystal structure and conformation of the synthetic cyclic tetrapeptide, cyclo(L -Pro-Sar)2, was determined by x-ray analysis. The peptide crystallizes in the orthorhombic space group P212121 with cell parameters a = 9.277(1), b = 12.884(1), and c = 15.581(2) Å. The crystal structure was solved by the symbolic addition procedure for direct phase determination and least-squares refinement using 1796 reflections, which led to the final R value of 0.043. This structure provides the first example observed in a crystal of a cyclic tetrapeptide in which all four peptide units have been found in the cis conformation with ω angles deviating slightly by 2°–10° from the ideal value of 0°. It was also found that the two Pro Cα-CO single bonds assumed a trans′ (ψ = 159.6° and 158.4°) conformation. Adjoining average planes of the peptide groups fall at nearly right angles to each other. The pyrrolidine ring conformations of the two prolyl residues are in the envelope form, with Cγ carbon out of the least-squares planes for the remaining four atoms.  相似文献   

6.
We prepared 125I-labeled cholera toxin B subunit (125I-labeled CT-B, specific activity 98 Ci/mmol) and found that its binding to rat IEC-6 intestinal epithelial cells was high-affinity (Kd 1.9 nM). The binding of labeled protein was completely inhibited by unlabeled thymosin-α1 (TM-α1), interferon-α2 (IFN-α2), and synthetic peptide LKEKK, which corresponds to residues 16–20 in TM-α1 and 131–135 in IFN-α2 (Ki 1.2, 0.9, and 1.6 nM, respectively), but was not inhibited by synthetic peptide KKEKL with inverted amino acid sequence (Ki > 10 μM). Thus, TM-α1, IFN-α2, and the LKEKK peptide bind with high affinity and specificity to CT-B receptor on rIEC-6 cells. It was found that CT-B and the LKEKK peptide at concentrations of 10–1000 nM increased nitric oxide production and soluble guanylate cyclase activity in the cells in a dose-dependent manner.  相似文献   

7.
Recently the folding of a staphylococcal nuclease (P117G) variant was examined with the hydrogen-deuterium (H-D) exchange technique. Many of the residues that showed significant protection are located in protection are located in β-sheet regions. About half the residues protected belong to an antiparallel β-hairpin structure (residues 21–35) in the native structure. The β-hairpin structure is formed by strands 2 and 3 of sheet 2 connected by the sequence27 Y KGQP31 in a type I′ reverse turn conformation with a 4 → 1 hydrogen bonding between Q30 NH and Y27 C=O. We have targeted the conformational characterization of the peptide model Ac-YKGQP-NH2 with 1II two-dimensional nmr techniques in aqueous solution with a view to assessing its propensity to sample turn conformational forms and thus initiate the formation of β-hairpin structure. Based upon the observed dαn (i, i + 1), dαn (i, i + 3), and dnn (i, i + 1) nuclear Overhauser effect connectivities, temperature coefficients for amide protons and conformational analysis with quantum mechanical perturbative configuration interaction over localized orbitals method, we conclude that the model peptide samples turn conformational forms with reduced conformational entropy. We suggest that the turn can nucleate the formation of the β-hairpin structure in the refolding of nuclease. Observation of turn propensity for this sequence is consistent with the folding mechanism of the Greek key motif (present in Staphylococcal nuclease) proposed in the literature. © 1997 John Wiley & Sons, Inc.  相似文献   

8.
Brevinin‐1BYa (FLPILASLAAKFGPKLFCLVTKKC), first isolated from skin secretions of the foothill yellow‐legged frog Rana boylii, shows broad‐spectrum activity, being particularly effective against opportunistic yeast pathogens. The structure of brevinin‐1BYa was investigated in various solution and membrane‐mimicking environments by proton nuclear magnetic resonance (1H‐NMR) spectroscopy and molecular modelling. The peptide does not possess a secondary structure in aqueous solution. In a 33% 2,2,2‐trifluoroethanol (TFE‐d3)‐H2O solvent mixture, as well as in membrane‐mimicking sodium dodecyl sulfate and dodecylphosphocholine micelles, the peptide's structure is characterised by a flexible helix‐hinge‐helix motif, with the hinge located at the Gly13/Pro14 residues, and the two α‐helices extending from Pro3 to Phe12 and from Pro14 to Thr21. Positional studies involving the peptide in sodium dodecyl sulfate and dodecylphosphocholine micelles using 5‐doxyl‐labelled stearic acid and manganese chloride paramagnetic probes show that the peptide's helical segments lie parallel to the micellar surface, with the residues on the hydrophobic face of the amphipathic helices facing towards the micelle core and the hydrophilic residues pointing outwards, suggesting that the peptide exerts its biological activity by a non–pore‐forming mechanism.  相似文献   

9.
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   

10.
A series of N- and C-protected, monodispersed homo-oligopeptides (to the dodecamer level) from the small-ring alicyclic Cα,α-dialkylated glycine 1-aminocyclobutane-1-carboxylic acid (Ac4c) and two Ala/Ac4c tripeptides were synthesized by solution methods and fully characterized. The conformational preferences of all the model peptides were determined in deuterochloroform solution by FT-IR absorption and 1H-NMR. The molecular structures of the amino acid derivatives Z-Ac4c-OH and Z2-Ac4c-OH, the tripeptides Z-(Ac4c)3-OtBu, Z-Ac4c-(L -Ala)2-OMe and Z-L -Ala-Ac4c-L -Ala-OMe, and the tetrapeptide Z-(Ac4c)4-OtBu were determined in the crystal state by X-ray diffraction. The average geometry of the cyclobutyl moiety of the Ac4c residue was assessed and the τ(N–Cα–C′) bond angle was found to be significantly expanded from the regular tetrahedral value. The conformational data are strongly in favour of the conclusion that the Ac4c residue is an effective β-turn and helix former. A comparison with the structural propensities of α-aminoisobutyric acid, the prototype of Cα,α-dialkylated glycines, and the other extensively investigated members of the family of 1-aminocycloalkane-1-carboxylic acids (Acnc, with n=3, 5–8) is made and the implications for the use of the Ac4c residue in conformationally constrained peptide analogues are briefly examined. © 1997 European Peptide Society and John Wiley & Sons, Ltd  相似文献   

11.
An α-amylase from Aspergillus oryzae, Taka-amylase A (TAA), was cleaved into peptide fragments by an acid protease. Inactivation of TAA was greatly retarded by the addition of α-cyclodextrin or Ca2+. TAA peptide fragments were separated into two groups having no and high affinity to the substrate, soluble starch. This separation was done by the forced affinity chromatography method by a column of epichlorohydrin cross-linked soluble starch gel. Three peptides were isolated from the high-affinity fragments, purified by the ODS-120T column, and their amino acids were sequenced. Peptides I, II, and III originated from α2-helix, α3-helix, and β2-sheet, respectively, and all of these were located in the (β/α)8 barrel of the main domain of TAA molecule. A stereo graphic view showed that Peptides I–III were at the cleft near the catalytic site. Occurrence of a Trp residue in all three peptides strongly suggested that Trp was very important in the binding of TAA to the substrate, soluble starch.  相似文献   

12.
The measurement of glycated hemoglobin A1c (HbA1c) has important implications for diagnosis of diabetes and assessment of treatment effectiveness. We proposed specific sequence motifs to identify enzymes that oxidize glycated compounds from genome database searches. The gene encoding a putative fructosyl amino acid oxidase was found in the Phaeosphaeria nodorum SN15 genome and successfully expressed in Escherichia coli. The recombinant protein (XP_001798711) was confirmed to be a novel fructosyl peptide oxidase (FPOX) with high specificity for α‐glycated compounds, such as HbA1c model compounds fructosyl‐αN‐valine (f‐αVal) and fructosyl‐αN‐valyl‐histidine (f‐αVal‐His). Unlike previously reported FPOXs, the P. nodorum FPOX has a Km value for f‐αVal‐His (0.185 mM) that is considerably lower than that for f‐αVal (0.458 mM). Based on amino acid sequence alignment, three dimensional structural modeling, and site‐directed mutagenesis, Gly60 was found to be a determining residue for the activity towards f‐αVal‐His. A flexible surface loop region was also found to likely play an important role in accepting f‐αVal‐His. Biotechnol. Bioeng. 2010; 106: 358–366. © 2010 Wiley Periodicals, Inc.  相似文献   

13.
Our previous studies of the potential utility of the CαDα stretch frequency, ν(CD), as a tool for determining conformation in peptide systems (Mirkin and Krimm, J Phys Chem A 2004, 108, 10923–10924; 2007, 111, 5300–5303) dealt with the spectroscopic characteristics of isolated alanine peptides with αR, β, and polyproline II structures. We have now extended these ab initio calculations to include various explicit‐water environments interacting with such conformers. We find that the structure‐discriminating feature of this technique is in fact enhanced as a result of the conformation‐specific interactions of the bonding waters, in part due to our finding (Mirkin and Krimm, J Phys Chem B 2008, 112, 15268) that Cα? Dα…O(water) hydrogen bonds can be present in addition to those expected between water and the CO and NH of the peptide groups. In fact, ν(CD) is hardly affected by the latter bonding but can be shifted by up to 70 cm?1 by the former hydrogen bonds. We also discuss the factors that will have to be considered in developing the molecular dynamics (MD) treatment needed to satisfactorily take account of the influence of outer water layers on the structure of the first‐layer water molecules that hydrogen bond to the peptide backbone. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 791–800, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

14.
We report proton magnetic resonance studies of a series of lysine oligopeptides in H2O solution. At pH 5 the protonated ε-amino groups are seen as broad resonances; the peptide NH proton resonances are split by spin–spin coupling with the Cα-H proton, and appear at positions which depend on position in the chain and on chain length. Assignments were made by the europium shift method, and we observed the expected effect of catalysis by the terminal —NH3+ of exchange of the adjacent peptide NH. Coupling constants and the temperature coefficient of chemical shift values were consistent with a non-hydrogen-bonded structure for the oligolysines. The rate and mechanism of NH hydrogen exchange were investigated by line-broadening measurements of the peptide protons as a function of pH. Exchange was found to be OH? catalyzed, with large differences in the rate depending on position in the chain. Preliminary studies of the complex between double-helical d(pA)3pGpC(pT)3 and tetra(L -lysine) were performed using 1H- and 31P-nmr techniques. Pmr spectra of the complex at pH values ranging from 3.98 to 8.15 showed very complicated patterns. Downfield shifts and reduction in exchange rates were observed for several tetra(L -lysine) protons. 31P-nmr spectra of the complex reveal an upfield shift of 1 ppm for 3′-5′ phosphate diester resonances on complexation. 31P T1 relaxation times change little on complex formation at low temperature but are altered at higher temperature.  相似文献   

15.
D A Torchia  C M Deber 《Biopolymers》1972,11(3):653-659
NMR spectra of cyclo(tri-L -prolyl), c-(P)3, show large shifts of the Hα resonance on adding C6D6 to a solution of c-(P)3 in CD2Cl2. CPK models and observed coupling constants indicate a rigid c-(P)3 conformation, independent of solvent composition, suggesting that these shifts result from formation of stereospecific C6D6–c-(P)3 collision complexes in which the c-(P)3 Hα lie near the face of the aromatic ring. The temperature dependence of the Hα shifts and the solvent dependent shifts observed on adding toluene-d8 or nitrobenzene-d5 to the c-(P)3 solution suggest that preferred C6D6–c-(P)3 orientations result from attractive interactions between the electron-rich aromatic ring and the electropositive Hα's and/or δ+ nitrogen atoms in the peptide backbone. Reports of such interactions in increasingly diverse peptide model systems suggest that they may play a role in stabilizing protein structures.  相似文献   

16.
Vpr, one of the accessory gene products encoded by HIV-1, is a 96-residue protein with a number of functions, including targeting of the viral pre-integration complex to the nucleus and inducing growth arrest of dividing cells. We have characterized by 2D NMR the solution conformations of bioactive synthetic peptide fragments of Vpr encompassing a pair of H(F/S)RIG sequence motifs (residues 71–75 and 78–82 of HIV-1 Vpr) that cause cell membrane permeabilization and death in yeast and mammalian cells. Due to limited solubility of the peptides in water, their structures were studied in aqueous trifluoroethanol. Peptide Vpr59–86 (residues 59–86 of Vpr) formed an α-helix encompassing residues 60–77, with a kink in the vicinity of residue 62. The first of the repeated sequence motifs (HFRIG) participated in the well-defined α-helical domain whereas the second (HSRIG) lay outside the helical domain and formed a reverse turn followed by a less ordered region. On the other hand, peptides Vpr71–82 and Vpr71–96, in which the sequence motifs were located at the N-terminus, were largely unstructured under similar conditions, as judged by their CαH chemical shifts. Thus, the HFRIG and HSRIG motifs adopt α-helical and turn structures, respectively, when preceded by a helical structure, but are largely unstructured in isolation. The implications of these findings for interpretation of the structure–function relationships of synthetic peptides containing these motifs are discussed. © 1998 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
An artificially inserted extra peptide (21 amino acid peptide) between the B. subtilis α-amylase signal peptide and the mature thermostable α-amylase was completely cleaved by B. subtilis alkaline protease in vitro. The cleavage to form a mature enzyme was observed between pH 7.5 and 10, but not between pH 6.0 and 6.5, although a similar protease activity toward Azocall was observed between pH 6.0 and 7.5. To analyze the effects of pH on the cleavage, CD spectra at pH 6, 8, and 11 of the NH2-terminally extended thermostable α-amylase were analyzed and the results were compared with those of the mature form of the α-amylase. It is suggesteded that the cleavage of the NH2-terminally extended peptide is controlled by the secondary and tertiary structure of the precursor enzyme. Similar cleavage of different NH2-terminally extended peptides by the alkaline protease was also found in other hybrid thermostable α-amylases obtained.  相似文献   

18.
Cleavage reactions at backbone loci are one of the consequences of oxidation of proteins and peptides. During α‐amidation, the Cα–N bond in the backbone is cleaved under formation of an N‐terminal peptide amide and a C‐terminal keto acyl peptide. On the basis of earlier works, a facilitation of α‐amidation by the thioether group of adjacent methionine side chains was proposed. This reaction was characterized by using benzoyl methionine and benzoyl alanyl methionine as peptide models. The decomposition of benzoylated amino acids (benzoyl‐methionine, benzoyl‐alanine, and benzoyl‐methionine sulfoxide) to benzamide in the presence of different carbohydrate compounds (reducing sugars, Amadori products, and reductones) was studied during incubation for up to 48 h at 80 °C in acetate‐buffered solution (pH 6.0). Small amounts of benzamide (0.3–1.5 mol%) were formed in the presence of all sugars and from all benzoylated species. However, benzamide formation was strongly enhanced, when benzoyl methionine was incubated in the presence of reductones and Amadori compounds (3.5–4.2 mol%). The reaction was found to be intramolecular, because α‐amidation of a similar 4‐methylbenzoylated amino acid was not enhanced in the presence of benzoyl‐methionine and carbohydrate compounds. In the peptide benzoyl‐alanyl‐methionine, α‐amidation at the methionine residue is preferred over α‐amidation at the benzoyl peptide bond. We propose here a mechanism for the enhancement of α‐amidation at methionine residues. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
The fully blocked pentapeptide Tfa-(Deg)2-L -Abu-(Deg)2-OtBu (Tfa:triflouroacetyl; Deg: Cα,α-diethylglycine; OtBu: tert-butoxy) adopts in the crystal state a regular, right-handed 310-helical structure stabilized by three N ? H …? O ? C intramolecular 1 ← 4 (or C10) H bonds, as determined by an x-ray diffraction analysis. However, a Fourier transform ir absorption and 1H-nmr study strongly supports the view that in deuterochloroform solution the four Deg residues at both termini of the peptide main chain are involved in successive, fully extended C5 forms. A comparison with the stable, fully developed, multiple C5 conformation of Tfa-(Deg)5-OtBu indicates that incorporation of an Abu guest residue, interrupting the side-chain uniformity of the host (Deg)5 homopeptide, while altering only marginally the conformation in a solvent of low polarity, is responsible for a dramatic perturbation of the crystal-state structure. © 1994 John Wiley & Sons, Inc.  相似文献   

20.
Protein–protein interactions play central roles in physiological and pathological processes. The bases of the mechanisms of drug action are relevant to the discovery of new therapeutic targets. This work focuses on understanding the interactions in protein–protein–ligands complexes, using proteins calmodulin (CaM), human calcium/calmodulin‐dependent 3′,5′‐cyclic nucleotide phosphodiesterase 1A active human (PDE1A), and myosin light chain kinase (MLCK) and ligands αII–spectrin peptide (αII–spec), and two inhibitors of CaM (chlorpromazine (CPZ) and malbrancheamide (MBC)). The interaction was monitored with a fluorescent biosensor of CaM (hCaM M124C–mBBr). The results showed changes in the affinity of CPZ and MBC depending on the CaM–protein complex under analysis. For the Ca2+–CaM, Ca2+–CaM–PDE1A, and Ca2+–CaM–MLCK complexes, CPZ apparent dissociation constants (Kds) were 1.11, 0.28, and 0.55 μM, respectively; and for MBC Kds were 1.43, 1.10, and 0.61 μM, respectively. In competition experiments the addition of calmodulin binding peptide 1 (αII–spec) to Ca2+hCaM M124C–mBBr quenched the fluorescence (Kd = 2.55 ± 1.75 pM) and the later addition of MBC (up to 16 μM) did not affect the fluorescent signal. Instead, the additions of αII–spec to a preformed Ca2+hCaM M124C–mBBr–MBC complex modified the fluorescent signal. However, MBC was able to displace the PDE1A and MLCK from its complex with Ca2+–CaM. In addition, docking studies were performed for all complexes with both ligands showing an excellent correlation with experimental data. These experiments may help to explain why in vivo many CaM drugs target prefer only a subset of the Ca2+–CaM regulated proteins and adds to the understanding of molecular interactions between protein complexes and small ligands. Copyright © 2013 John Wiley & Sons, Ltd.  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