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 62 毫秒
1.
Rabbit muscle troponin C was selectively modified at Cys-98 by 1,3-difluoro-4,6-dinitrobenzene. The second function of the bifunctional reagent was triggered at alkaline pH in the presence and absence of Ca2+. The crosslinked troponin C was hydrolyzed by trypsin and the peptides containing a dinitrobenzene moiety were isolated. When troponin C was crosslinked in the presence of Ca2+, the single dinitrobenzene-containing peptide was Gly-89-Arg-100, in which Cys-98 was crosslinked with Lys-90. When crosslinking was performed in the absence of Ca2+, beside the above peptide two additional peptides containing dinitrobenzene were found. One of these peptides is made up of two fragments, Ser-91-Arg-100 and Asn-105-Arg-120, crosslinked between Cys-98 and Tyr-109. The second peptide, Ala-121-Lys-140, contains modified Lys-136, presumably crosslinked with His-135. The data indicate that the distances between the α-carbon of Cys-98 and those of Lys-90, Tyr-109, Lys-136 and probably the α-carbon distance His-125-Lys-136, do not exceed 14 Å. Comparison with the X-ray structure of troponin C (Herzberg, O, and James, M.N.G. (1985) Nature 313, 653–659) indicates that some of the above distances increase on Ca2+-binding.  相似文献   

2.
Tritium exchange was used as a probe of transfer RNA structure in experiments with unfractionated tRNA (tRNAUnfrac and homogeneous tRNA3Leu from bakers' yeast. Exchange kinetics were measured over a range of ionic conditions that vary in ability to stabilize the secondary and tertiary structure of tRNA. The native conformations of both samples show the same kinetics of exchange. The kinetics for tRNA3Leu trapped in a denatured state in a “native” solvent are much faster, reflecting the conformation and not the ionic medium. In 0.1 M-Na+, where tRNA3Leu is denatured, the kinetics for tRNAUnfrac are intermediate between those for native and denatured tRNA3Leu, suggesting that in this solvent at 0 °C some tRNAs are denatured whereas other are still native. Upon further lowering of Na+ concentration, tRNAUnfrac shows increasingly faster exchange, suggesting complete electrostatic denaturation of the tertiary structure of all the tRNAs in the sample, and even disruption of secondary structure.Extrapolation of the essentially linear early-time kinetics to zero time provides minimal estimates of the number of slowly exchanging hydrogens. For native tRNA3Leu the number is 111±2 hydrogens, whereas for the trapped denatured conformation it is only 95±2. This difference reflects a smaller number of hydrogen-bonded bases in the denatured conformation. In 1 M-Na+, 101±2 slowly exchanging hydrogens are found for the native tRNA3Leu conformation, suggesting an incompletely formed native structure. For native tRNAUnfrac the comparable number is 101±3. These numbers of slowly exchanging hydrogens in the native conformations are consistent with tertiary structural hydrogen-bonding. Furthermore, this tertiary structure must be responsible for the slower exchange by native tRNA. The observed numbers of exchangeable hydrogens provide a basis for comparison of hydrogen-bonding interactions in native and denatured tRNA conformations.The mechanism of renaturation was also investigated, using tritium exchange as a monitor of perturbation of base pairing during the transition. When tRNAUnfrac in low Na+ is renatured by addition of Mg2+ during tritium exchangeout, a burst of exchange or “spillage” of tritium is detected. This suggests that a fraction of the base pairs of the rapidly renaturing tRNAs in the mixture is disrupted during renaturation. In that event, and by analogy with tRNA3Leu, part of the base-pairing arrangement of the denatured conformations may not be preserved in the native state; and if the native conformation includes the full “cloverleaf” pattern of secondary structure, that pattern may not be intact in some denatured conformations.  相似文献   

3.
ProP, an osmoprotectant symporter from the major facilitator superfamily was expressed, purified and reconstituted into proteoliposomes that are amenable to structural characterization using infrared spectroscopy. Infrared spectra recorded in both 1H2O and 2H2O buffers reveal amide I band shapes that are characteristic of a predominantly α-helical protein, and that are similar to those recorded from the well-characterized homolog, lactose permease (LacY). Curve-fit analysis shows that ProP and LacY both exhibit a high α-helical content. Both proteins undergo extensive peptide hydrogen-deuterium exchange after exposure to 2H2O, but are surprisingly thermally stable with denaturation temperatures greater than 60 °C. 25-30% of the peptide hydrogens in both ProP and LacY are resistant to exchange after 72 h in 2H2O at 4 °C. Surprisingly, these exchange resistant peptide hydrogens exchange completely for deuterium at temperatures below those that lead to denaturation. Our results show that ProP adopts a highly α-helical fold similar to that of LacY, and that both transmembrane folds exhibit unusually high temperature-sensitive solvent accessibility. The results provide direct evidence that ProP adopts a structure consistent with other major facilitator superfamily members.  相似文献   

4.
Amyloid deposits of amylin in the pancreas are an important characteristic feature found in patients with Type-2 diabetes. The aggregate has been considered important in the disease pathology and has been studied extensively. However, the secondary structures of the individual peptide have not been clearly identified. In this work, we present detailed solution structures of rat amylin using a combination of Monte Carlo and molecular dynamics simulations. A new Monte Carlo method is presented to determine the free energy of distinct biomolecular conformations. Both folded and random-coil conformations of rat amylin are observed in water and their relative stability is examined in detail. The former contains an α-helical segment comprised of residues 7-17. We find that at room temperature the folded structure is more stable, whereas at higher temperatures the random-coil structure predominates. From the configurations and weights we calculate the α-carbon NMR chemical shifts, with results that are in reasonable agreement with experiments of others. We also calculate the infrared spectrum in the amide I stretch regime, and the results are in fair agreement with the experimental line shape presented herein.  相似文献   

5.
Syncatalytic inactivation of pig heart cytoplasmic aspartate aminotransferase by β-chloro-[U-14C]L-alanine resulted in the incorporation of radioactivity corresponding to one mole of the label per mole of the monomeric unit of the enzyme. A borohydride-reduced and then carboxymethylated preparation of the labeled enzyme was digested by trypsin. A radioactive peptide was isolated and found to contain a covalently linked pyridoxyl derivative which absorbed at 325 nm. The amino acid sequence of this peptide was Tyr-Phe-Val-Ser-Glu-Gly-Phe -Glu-Leu-Phe-Cys-Ala-Gln-Ser-Phe-Ser-Lys-Asn-Phe-Gly-Leu-Tyr-Asn-Glu-Arg. In the peptide the phosphopyridoxyl group seems to be covalently bound via alanyl moiety derived from β-chloro-L-alanine, the β-carbon atom of which is covalently linked to the ?-nitrogen atom of the lysyl residue(Lys). From a comparison with the amino acid composition of the phosphopyridoxyl peptide isolated from the tryptic digest of a borohydride-reduced holoenzyme, it was concluded that the modified lysul residue was identical to that involved in binding pyridoxal phosphate to the apoenzyme.  相似文献   

6.
Here we present a novel NMR method for the structure determination of calcium-calmodulin (Ca2+-CaM)-peptide complexes from a limited set of experimental restraints. A comparison of solved CaM-peptide structures reveals invariability in CaM’s backbone conformation and a structural plasticity in CaM’s domain orientation enabled by a flexible linker. Knowing this, the collection and analysis of an extensive set of NOESY spectra is redundant. Although RDCs can define CaM domain orientation in the complex, they lack the translational information required to position the domains on the bound peptide and highlight the necessity of intermolecular NOEs. Here we employ a specific isotope labeling strategy in which the role of methionine in CaM-peptide interactions is exploited to collect these critical NOEs. By 1H, 13C-labeling the methyl groups of deuterated methionine against a 2H, 12C background, we can acquire a 13C-edited NOESY characterized by simplified, easily analyzable spectra. Together with measured CaM backbone HN-N RDCs and intrapeptide NOE-based distances, these intermolecular NOEs provide restraints for a low temperature torsion-angle dynamics and simulated annealing protocol used to calculate the complex structure. We have applied our method to a CaM complex previously solved through X-ray crystallography: Ca2+-CaM bound to the CaM kinase I peptide (PDB code: 1MXE). The resulting structure has a backbone RMSD of 1.6 Å to that previously published. We have also used this test complex to investigate the importance of homologous model selection on the calculated outcome. In addition to having application for fast complex structure determination, this method can be used to determine the structures of difficult complexes characterized by chemical shift overlap and broad signals for which the traditional method based on the use of fully 13C, 15N-labeled CaM fails.  相似文献   

7.
The equilibrium between the cis and trans forms of X-Pro peptide bonds can readily be measured in the 13C nmr spectra. In the present paper we investigate how observation of this equilibrium could be used as an nmr probe for conformational studies of flexible polypeptide chains. The experiments include studies by 13C nmr of a series of linear oligopeptides containing different X-L -Pro peptide bonds, with X = Gly, L -Ala, L -Leu, L -Phe, D -Ala, D -Leu, and D -Phe. Overall the study confirms that X-Pro peptide bonds can generally be useful as 13C nmr probes reporting the formation of nonrandom conformation in flexible polypeptide chains. It was found that the cistrans equilibrium of X-Pro is greatly affected by the side chain of X and the configuration of the α-carbon atom of X. On the basis of these observations some general rules are suggested for a practical applications of the X-Pro nmr probes in conformational studies of polypeptide chains.  相似文献   

8.
In order to investigate the importance of the monomeric γ-carbon chemistry in lignin biopolymerization and structure, synthetic lignins (dehydrogenation polymers; DHP) were made from monomers with different degrees of oxidation at the γ-carbon, i.e., carboxylic acid, aldehyde and alcohol. All monomers formed a polymeric material through enzymatic oxidation. The polymers displayed similar sizes by size exclusion chromatography analyses, but also exhibited some physical and chemical differences. The DHP made of coniferaldehyde had poorer solubility properties than the other DHPs, and through contact angle of water measurement on spin-coated surfaces of the polymeric materials, the DHPs made of coniferaldehyde and carboxylic ferulic acid exhibited higher hydrophobicity than the coniferyl alcohol DHP. A structural characterization with 13C NMR revealed major differences between the coniferyl alcohol-based polymer and the coniferaldehyde/ferulic acid polymers, such as the predominance of aliphatic double bonds and the lack of certain benzylic structures in the latter cases. The biological role of the reduction at the γ-carbon during monolignol biosynthesis with regard to lignin polymerization is discussed.  相似文献   

9.
The effects of Cu2+ binding and the utilization of different force fields when modeling the structural characteristics of α-syn12 peptide were investigated. To this end, we performed extensive temperature replica exchange molecular dynamics (T-REMD) simulations on Cu2+-bound and unbound α-syn12 peptide using the GROMOS 43A1, OPLS-AA, and AMBER03 force fields. Each replica was run for 300 ns. The structural characteristics of α-syn12 peptide were studied based on backbone dihedral angle distributions, free-energy surfaces obtained with different reaction coordinates, favored conformations, the formation of different Turn structures, and the solvent exposure of the hydrophobic residues. The findings show that AMBER03 prefers to sample helical structures for the unbound α-syn12 peptide and does not sample any β-hairpin structure for the Cu2+-bound α-syn12 peptide. In contrast, the central structure of the major conformational clusters for the Cu2+-bound and unbound α-syn12 peptide according to simulations performed using the GROMOS 43A1 and OPLS-AA force fields is a β-hairpin with Turn9-6. Cu2+ can also promote the formation of the β-hairpin and increase the solvent exposure of hydrophobic residues, which promotes the aggregation of α-syn12 peptide. This study can help us to understand the mechanisms through which Cu2+ participates in the fibrillation of α-syn12 peptide at the atomic level, which in turn represents a step towards elucidating the nosogenesis of Parkinson’s disease.
Figure
The representative structures of Cu2+-bound and unbound α-syn12 peptide using three different force fields  相似文献   

10.
11.
Using amide hydrogen exchange combined with electrospray ionization mass spectrometry, we have in this study determined the number of amide hydrogens on several peptides that become solvent-inaccessible as a result of their high-affinity interaction with the urokinase-type plasminogen activator receptor (uPAR). These experiments reveal that at least six out of eight amide hydrogens in a synthetic nine-mer peptide antagonist (AE105) become sequestered upon engagement in uPAR binding. Various uPAR mutants with decreased affinity for this peptide antagonist gave similar results, thereby indicating that deletion of the favorable interactions involving the side chains of these residues in uPAR does not affect the number of hydrogen bonds established by the main chain of the peptide ligand. The isolated growth factor-like domain (GFD) of the cognate serine protease ligand for uPAR showed 11 protected amide hydrogens in the receptor complex. Interestingly, a naturally occurring O-linked fucose on Thr(18) confers protection of two additional amide hydrogens in GFD when it forms a complex with uPAR. Dissociation of the uPAR-peptide complexes is accompanied by a correlated exchange of nearly all amide hydrogens on the peptide ligand. This yields bimodal isotope patterns from which dissociation rate constants can be determined. In addition, the distinct bimodal isotope distributions also allow investigation of the exchange kinetics of receptor-bound peptides providing information about the local structural motions at the interface. These exchange experiments therefore provide both structural and kinetic information on the interaction between uPAR and these small peptide antagonists, which in model systems show promise as inhibitors of intravasation of human cancer cells.  相似文献   

12.
Ten C-glycosyl β2- and β/β2-peptides with three to eight amino acid residues have been prepared. Solution and solid-phase peptide syntheses were employed to assemble β2-amino acids in which C-glycosylic substituents are attached to the C-2 position of β-amino acids. Conformational analysis of the C-glycosyl β2-peptides using NMR and CD spectra indicates that the tripeptide can form a helical secondary structure. Besides, helix directions of the C-glycosyl β/β2-peptides are governed by the configuration at the α-carbon of the peptide backbone that originates from the stereocenter of the C-glycosyl β2-amino acids.  相似文献   

13.
Interconversion of d-ribose-5-phosphate (R5P) and d-ribulose-5-phosphate is an important step in the pentose phosphate pathway. Two unrelated enzymes with R5P isomerase activity were first identified in Escherichia coli, RpiA and RpiB. In this organism, the essential 5-carbon sugars were thought to be processed by RpiA, while the primary role of RpiB was suggested to instead be interconversion of the rare 6-carbon sugars d-allose-6-phosphate (All6P) and d-allulose-6-phosphate. In Mycobacterium tuberculosis, where only an RpiB is found, the 5-carbon sugars are believed to be the enzyme's primary substrates. Here, we present kinetic studies examining the All6P isomerase activity of the RpiBs from these two organisms and show that only the E. coli enzyme can catalyze the reaction efficiently. All6P instead acts as an inhibitor of the M. tuberculosis enzyme in its action on R5P. X-ray studies of the M. tuberculosis enzyme co-crystallized with All6P and 5-deoxy-5-phospho-d-ribonohydroxamate (an inhibitor designed to mimic the 6-carbon sugar) and comparison with the E. coli enzyme's structure allowed us to identify differences in the active sites that explain the kinetic results. Two other structures, that of a mutant E. coli RpiB in which histidine 99 was changed to asparagine and that of wild-type M. tuberculosis enzyme, both co-crystallized with the substrate ribose-5-phosphate, shed additional light on the reaction mechanism of RpiBs generally.  相似文献   

14.
To study the catalytic mechanism of phosphorylation catalyzed by cAMP-dependent protein kinase (PKA) a structure of the enzyme-substrate complex representing the Michaelis complex is of specific interest as it can shed light on the structure of the transition state. However, all previous crystal structures of the Michaelis complex mimics of the PKA catalytic subunit (PKAc) were obtained with either peptide inhibitors or ATP analogs. Here we utilized Ca2+ ions and sulfur in place of the nucleophilic oxygen in a 20-residue pseudo-substrate peptide (CP20) and ATP to produce a close mimic of the Michaelis complex. In the ternary reactant complex, the thiol group of Cys-21 of the peptide is facing Asp-166 and the sulfur atom is positioned for an in-line phosphoryl transfer. Replacement of Ca2+ cations with Mg2+ ions resulted in a complex with trapped products of ATP hydrolysis: phosphate ion and ADP. The present structural results in combination with the previously reported structures of the transition state mimic and phosphorylated product complexes complete the snapshots of the phosphoryl transfer reaction by PKAc, providing us with the most thorough picture of the catalytic mechanism to date.  相似文献   

15.
Possible effects of changes in net charge on protein hydrogen exchange rates were investigated by desalting hen egg-white lysozyme, which allowed its net charge to increase with decreasing pH in the acid region. Chloride ion-binding ratios, expressed as ratios of free to total Cl?, were measured with a chloride-specific electrode at pH 5 on a 2.4% solution of a five-time-desalted product. This ratio was used to show a 97% reduction of the 11% Cl? present in a commercial lysozyme preparation upon three passes of the enzyme through a column of ion-retardation resin. Net charges on the purified product were assigned from a combination of electrophoretic mobility and proton titration data gathered under minimal ionic strength conditions. The net charge on the desalted product increased by 1.64 units between pH 5.0 and 3.0. Hydrogendeuterium exchange studies on the purified lysozyme in D2O were obtained using the near-infrared region of a Cary 14R spectrophotometer. The rate-pD profile for k2, the rate constant for the intermediate class of exchanging hydrogens, showed a decrease in the apparent pD of minimum exchange rate of 0.3 units, when compared to that obtained earlier in 0.2 m added NaCl. However, the rate of exchange at pD minimum and the number of hydrogens in the class remained largely unaffected. A similar shift was observed for the rate-pD profile of the class 1 hydrogens. Thus, the effect of an increase in net positive charge is to shift the rate-pD profile to a lower pD. Moreover, the effect extended to the interior peptide hydrogens of this globular protein. Consequently, the exchange rates of all the observable hydrogens are altered by the net charge changes, and the effect appeared uniform. The shift can be accounted for quantitatively by applying electrostatic interaction terms to the acid and base catalytic constants characterizing the exchange process. The calculated electrostatic interaction factors in minimal salt and 0.2 m added NaCl were found to be 29 and 18% lower, respectively, than those obtained theoretically. Therefore, under conditions where changes in net charge may occur for a globular protein, the effect on hydrogen exchange rates can be estimated fairly well theoretically, especially at moderate ionic strengths.  相似文献   

16.
Conformational analyses of cyclic tetrapeptides consisting of alternating cis and trans peptide units have been made using contact criteria and energy calculations. This study has been restricted to those structures having a symmetry element in the backbone ring, such as a twofold axis (d) or a center of inversion (i). There are five main results. (1) There are two distinct types of conformations, which are stereochemically favorable corresponding to each of twofold and inversion-symmetrical structures, designated as d1, d2 (for twofold symmetrical) and i1, i2 (for inversion-symmetrical). Among these, the i1 type has the lowest energy when glycyl residues occur at all four α-carbon atoms. (2) With the glycyl residue at all four α-carbon atoms, methyl substitution at the cis peptide nitrogen atoms is possible in all the four types, whereas the substitution at trans peptide nitrogen atoms is possible only for the i1 type. Thus only in the i1 type can all the nitrogen atoms be methylated simultaneously. The conformation of the molecule in the crystal structure of cyclotetrasarcosyl belongs to the i1 type. (3) When alanyl residues occur at all four α-carbon atoms, the possible symmetrical type is dependent on the enantiomorphic form and the actual sequence of the alanyl residues. (4) The methyl substitution at peptide nitrogen atoms for cyclic tetrapeptides having alanyl residues causes more stereochemical restriction in the allowed conformations than with glycyl residues. (5) The prolyl residue can be incorporated favorably at the cis-trans junction of both d and i types of structures. The results of the present study are compared with the data on cyclic tetrapeptides available from the crystal structure and nmr studies. The results show an overall agreement both regarding the type of symmetry and the conformational parameters.  相似文献   

17.
S Segawa  K Kume 《Biopolymers》1986,25(10):1981-1996
The hydrogen-exchange reactions of peptide NH groups in lysozyme were studied by the change in the intensity of the amide II band in the ir spectrum. The slowest exchanging hydrogens, which are involved in intramolecular hydrogen bonding, are further divided into two groups at lower temperatures; half of them are exchanged through local unfolding and the other half through major cooperative unfolding. In order to study the correlation of the change in hydrogen-exchange rates with the change in the unfolding rate constant, we observed the effects of intrachain cross-linking, the addition of denaturant and ligand binding on the exchange rates through local unfolding. Although the exchange rate through major unfolding is greatly decreased by intrachain cross-linking between Glu 35 and Trp 108 (1/22000), the exchange rate through local unfolding is only slightly decreased (1/20). Even at higher temperatures, where most intact lysozyme molecules unfold, the folded conformation of cross-linked lysozyme remains compact, and no intermediate exists in which many side-chain atoms are packed loosely so that the hydrogen-exchange reaction occurs rapidly. Neither the addition of 2-PrOD molecules nor (NAG)3 binding affects the exchange rates through local unfolding. Our experiments confirm that the change in the unfolding rate constant does not correlate with the change in fluctuations in the relatively flexible hydrogen-bonded structure through which the exchange of peptide hydrogens takes place.  相似文献   

18.
Over the last decades the field of pharmaceutically relevant peptides has enormously expanded. Among them, several peptide families exist that contain three or more disulfide bonds. In this context, elucidation of the disulfide patterns is extremely important as these motifs are often prerequisites for folding, stability, and activity. An example of this structure-determining pattern is a cystine knot which comprises three constrained disulfide bonds and represents a core element in a vast number of mechanically interlocked peptidic structures possessing different biological activities. Herein, we present our studies on disulfide pattern determination and structure elucidation of cystine-knot miniproteins derived from Momordica cochinchinensis peptide MCoTI-II, which act as potent inhibitors of human matriptase-1. A top-down mass spectrometric analysis of the oxidised and bioactive peptides is described. Following the detailed sequencing of the peptide backbone, interpretation of the MS3 spectra allowed for the verification of the knotted topology of the examined miniproteins. Moreover, we found that the fragmentation pattern depends on the knottin’s folding state, hence, tertiary structure, which to our knowledge has not been described for a top-down MS approach before.  相似文献   

19.
Information was obtained on rates of overall molecular reorientation and segmental motion of amino acid sidechains of oxytocin in dimethylsulfoxide by determination of spin-lattice relaxation times (T1) at 25 MHz for carbon-13 in natural abundance in the hormone. The T1 values of the α-carbons of amino acid residues located in the 20-membered ring of oxytocin are all about 50 msec. The overall correlation time for the hormone backbone was estimated to be 8.8 × 10?10 sec. The sidechains of Tyr, Ile and Gln undergo segmental motion with respect to the backbone of the ring. The T1 value of the α-carbon of the Leu residue is greater than for any α-carbon in the ring, indicating an increased mobility of the backbone of the C-terminal acyclic peptide as compared to the ring. The β- and γ-carbons of the Pro residue undergo an exo-endo interconversion with regard to the plane formed by α-carbon, δ-carbon and N atom of the Pro pyrollidine ring. These data are discussed in light of results from other experimental and theoretical studies, including carbon-13 spin-lattice relaxation times for oxytocin in aqueous solution.  相似文献   

20.
The N-terminal domain of human immunodeficiency virus (HIV)-1 glycoprotein 41,000 (FP; residues 1–23; NH2-AVGIGALFLGFLGAAGSTMGARS-CONH2) is involved in the fusion and cytolytic processes underlying viral-cell infection. Here, we use circular dichroism (CD) spectroscopy, along with electrospray ionization (ESI) mass spectrometry and tandem (MS/MS) mass spectrometry during the course of hydrogen/deuterium exchange, to probe the local conformations of this synthetic peptide in two membrane mimics. Since amino acids that participate in defined secondary structure (i.e., α-helix or β-sheet) exchange amido hydrogens more slowly than residues in random structures, deuterium exchange was combined with CD spectroscopy to map conformations to specific residues. For FP suspended in the highly structure-promoting solvent hexafluoroisopropanol (HFIP), CD spectra indicated high α-helix and disordered structures, whereas ESI and MS/MS mass spectrometry indicated that residues 5–15 were α-helical and 16–23 were disordered. For FP suspended in the less structure-promoting solvent trifluoroethanol (TFE), CD spectra showed lower α-helix, with ESI and MS/MS mass spectrometry indicating that only residues 9–15 participated in the α-helix. These results compare favorably with previous two-dimensional nuclear magnetic resonance studies on the same peptide. Proteins Suppl. 2:38–49, 1998. © 1998 Wiley-Liss, 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