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.
《Free radical research》2013,47(1-2):95-101
Spin-label nitroxyl derivatives of tetramethylpyrroline and tetramethylpyrrolidine in frozen solutions of perdeuterated methanol have been characterized by electron nucleus double resonance (ENDOR spec-troscopy). With use of selectively deuterated derivatives of 2,2,5,5-tetramethylpyrroline-l-oxyl-3-carboxamide, proton ENDOR resonance features have been assigned to the vinylic proton in the five membered pyrrolinyl ring and to the methyl groups. The ENDOR resonance features were analyzed on the basis of their dependence on H0. Two pairs of resonance features were assigned to the vinylic proton and were shown to correspond to ‖ and ⊥ hyperfine coupling (hfc) components. Six pairs of resonance features were ascribed to the methyl groups. The proton ENDOR spectra of the 3-carboxylic acid spin-label derivatives of tetramethylpyrroline and of tetramethylpyrrolidine compounds exhibited comparable features with nearly identical line splittings. From the observed ENDOR splittings, we have estimated the isotropic hfc component of the vinylic proton in 2,2,5,5-tetramethylpyrroline-l-oxyl-3-carboxamide to be -1.81 ± 0.04 MHz in frozen methanol. On the basis of the anisotropic dipolar hfc components, the electron-to-vinylic proton distance is estimated as 3.78 ± 0.01 Å. in excellent agreement with that of 3.79 Å calculated from X-ray defined coordinates.  相似文献   

2.
The kinetics of the reaction between the stable free radical 1,1-diphenyl-2-picrylhydrazyl (DPPH) and methylated urates was studied. Urates that had methyl groups on the 1,3,9, or on the 1 and 3 or 1 and 9 nitrogens reacted with DPPH 15 to 77% faster than uric acid. Urates substituted with methyl groups on the 7 nitrogen or on both the 3 and 9 nitrogens reacted with DPPH at rates that were less than 0.1 that of uric acid. 3,7,9-Trimethyluric acid and 1,3,7,9-tetramethyluric acid reacted with DPPH at barely detectable rates. DPPH reacted with uric acid, the monomethylated urates, and some of the dimethylated urates in a ratio of 2:1. DPPH reacted with other dimethylated and trimethylated urates in a ratio of 1:1. Semiempirical MNDO calculations indicate that the most stable radical of uric acid is formed by hydrogen abstraction from the 3, 7 or 9 position. The most stable species resulting from loss of a second hydrogen lack hydrogens at the 3 and 7 positions or the 7 and 9 positions. For maximum reactivity with DPPH, methylated uric acid derivatives must have a hydrogen at nitrogen 7 and one of the hydrogens at either the 3 or 9 position.  相似文献   

3.
Single crystals of the four aromatic bioamine salts phenylethylamine hydrochloride, tyramine hydrochloride, tryptamine hydrochloride, and histamine dihydrochloride were grown in various states of deuteration. Free radicals were produced by exposure to X-rays between 77 and 300 K and investigated by electron spin resonance spectroscopy. Dissociation of atomic hydrogen from C beta of the aliphatic chain occurs in all compounds studied except tryptamine. However deamination as usually present in the analogous amino acids is not found. The C beta-radical is characterized by an anisotropic H alpha-splitting and two isotropic H beta-splittings. The latter splittings depend strongly on temperature in tyramine. In comparison to the analogous amino acids, radical formation in the aromatic residues is favoured. Among the seven different aromatic radicals found only one is identified in histamine but two in each of the three other bioamines. Two of these are characterized by hydrogen dissociation which occurs in phenylethylamine and tyramine. One hydrogen addition radical is found in each of the three compounds phenylethylamine, tyramine and histamine. In tryptamine two different addition radicals are detected. One of the two products can be converted into the other by visible light. The reverse process is induced by heat, thus permitting the switching of the radical site reversibly between two different structures.  相似文献   

4.
During porphyrin biosynthesis the oxygen-independent coproporphyrinogen III oxidase (HemN) catalyzes the oxidative decarboxylation of the propionate side chains of rings A and B of coproporphyrinogen III to form protoporphyrinogen IX. The enzyme utilizes a 5'-deoxyadenosyl radical to initiate the decarboxylation reaction, and it has been proposed that this occurs by stereo-specific abstraction of the pro-S-hydrogen atom at the beta-position of the propionate side chains leading to a substrate radical. Here we provide EPR-spectroscopic evidence for intermediacy of the latter radical by observation of an organic radical EPR signal in reduced HemN upon addition of S-adenosyl-L-methionine and the substrate coproporphyrinogen III. This signal (g(av) = 2.0029) shows a complex pattern of well resolved hyperfine splittings from at least five different hydrogen atoms. The radical was characterized using regiospecifically labeled (deuterium or 15N) coproporphyrinogen III molecules. They had been generated from a multienzyme mixture and served as efficient substrates. Reaction of HemN with coproporphyrinogen III, perdeuterated except for the methyl groups, led to the complete loss of resolved proton hyperfine splittings. Substrates in which the hydrogens at both alpha- and beta-positions, or only at the beta-positions of the propionate side chains, or those of the methylene bridges, were deuterated showed that there is coupling with hydrogens at the alpha-, beta-, and methylene bridge positions. Deuterium or 15N labeling of the pyrrole nitrogens without labeling the side chains only led to a slight sharpening of the radical signal. Together, these observations clearly identified the radical signal as substrate-derived and indicated that, upon abstraction of the pro-S-hydrogen atom at the beta-position of the propionate side chain by the 5'-deoxyadenosyl radical, a comparatively stable delocalized substrate radical intermediate is formed in the absence of electron acceptors. The observed hyperfine constants and g values show that this coproporphyrinogenyl radical is allylic and encompasses carbon atoms 3', 3, and 4.  相似文献   

5.
The proposal that EPR Signal II in spinach chloroplasts is due to a plastoquinone cation radical (O'Malley, P.J. and Babcock, G.T. (1983) Biophys. J. 41, 315a) has been investigated in further detail. The similarity in spectral shape between Signal II and the 2-methyl-5-isopropylhydroquinone cation radical is shown to arise from hyperfine coupling to one methyl group for both radicals. A well-resolved four line EPR spectrum of approximate relative intensity 1:3:3:1 for membrane orientation parallel and perpendicular to the applied magnetic field direction also indicates that the partially resolved structure of Signal II is due to hyperfine interaction with one methyl group, i.e., the 2-CH3 group of the plastoquinone cation radical. The ENDOR band observed for this coupling is similar to that observed for methyl group bands of model quinone radicals. The principal hyperfine tensor values obtained for the methyl group interactions are A = 27.2 MHz and A = 31.4 MHz. The large isotropic coupling value (28.6 MHz) of the plastoquinone cation radical's 2-methyl group in vivo indicates that the antisymmetric orbital is the sole contributor to the spin-density distribution of Signal II. The orientation data also suggest that the plastoquinone cation radical is oriented such that the C-CH3 bond direction, and hence the aromatic ring plane, lies perpendicular to the membrane plane.  相似文献   

6.
Warncke K 《Biochemistry》2005,44(9):3184-3193
Molecular structural features of the product radical in the Co(II)-product radical pair catalytic intermediate state in coenzyme B(12)- (adenosylcobalamin-) dependent ethanolamine deaminase from Salmonella typhimurium have been characterized by using X-band three-pulse electron spin-echo envelope modulation (ESEEM) spectroscopy in the disordered solid state. The Co(II)-product radical pair state was prepared by cryotrapping holoenzyme during steady-state turnover on excess 1,1,2,2-(2)H(4)-aminoethanol or natural abundance, (1)H(4)-aminoethanol. Simulation of the (2)H/(1)H quotient ESEEM (obtained at two microwave frequencies, 8.9 and 10.9 GHz) from the interaction of the unpaired electron localized at C2 of the product radical with nearby (2)H nuclei requires four types of coupled (2)H, which are assigned as follows: (a) a single strongly coupled (effective dipole distance, r(eff) = 2.3 A) (2)H in the C5' methyl group of 5'-deoxyadenosine, (b) two weakly coupled (r(eff) = 4.2 A) (2)H in the C5' methyl group, (c) one (2)H coupling from a beta-(2)H bonded to C1 of the product radical (isotropic hyperfine coupling, A(iso) = 4.7 MHz), and (d) a second type of C1 beta-(2)H coupling (A(iso) = 7.7 MHz). The two beta-(2)H couplings are proposed to arise from two C1-C2 rotamer states of the product radical that are present in approximately equal proportion. A model is presented, in which C5' is positioned at a distance of 3.3 A from C2, which is comparable with the C1-C5' distance in the Co(II)-substrate radical pair intermediate. Therefore, the C5'methyl group remains in close (van der Waals) contact with the substrate and product radical species during the radical rearrangement step of the catalytic cycle, and the C5' center is the sole mediator of radical pair recombination in ethanolamine deaminase.  相似文献   

7.
Peloquin JM  Tang XS  Diner BA  Britt RD 《Biochemistry》1999,38(7):2057-2067
We have used electron spin-echo envelope modulation spectroscopy (ESEEM) to characterize the protein-cofactor interactions present in the QA- binding pocket of PS II centers isolated from spinach and Synechocystis. We conclude that the ESEEM spectrum of QA- is the result of interactions of the S = 1/2 electron spin of QA- with the I = 1 nuclear spins of the peptide nitrogens of two different amino acids. One peptide nitrogen has ESEEM peaks near 0.7, 2.0, 2.85, and 5.0 MHz with isotropic and dipolar hyperfine couplings of Aiso = 2.0 MHz and Adip = 0.25 MHz, respectively. On the basis of these hyperfine couplings we predict the existence of a strong hydrogen bond between QA- and the peptide nitrogen with a hydrogen bond distance of about 2 A. We have not identified the amino acid origin of this peptide nitrogen. By using amino acid specific isotopic labeling in conjunction with site-directed mutagenesis, we demonstrate that the second peptide nitrogen is that of D2-Ala260, with ESEEM peaks near 0.6 and 1.5 MHz and an isotropic hyperfine coupling, Aiso, less than 0.2 MHz. This small isotropic coupling suggests that the D2-Ala260 peptide nitrogen at best forms a weak hydrogen bond with QA-.  相似文献   

8.
The fast internal dynamics of human ubiquitin have been studied by the analysis of 15N relaxation of backbone amide nitrogens. The amide 15N resonances have been assigned by use of heteronuclear multiple-quantum spectroscopy. Spin lattice relaxation times at 60.8 and 30.4 MHz and the steady-state nuclear Overhauser effect at 60.8 MHz have been determined for 67 amide 15N sites in the protein using two-dimensional spectroscopy. These data have been analyzed in terms of the model free treatment of Lipari and Szabo [Lipari, G., & Szabo, A. (1982) J. Am. Chem. Soc. 104, 4546-4559]. The global motion of the protein is shown to be isotropic and is characterized by a correlation time of 4.1 ns rad-1. The generalized order parameters (S2) of backbone amide N-H vectors in the globular region of the protein range from 0.5 to 0.95. No apparent correlation between secondary structure and generalized order parameters is observed. There is, however, a strong correlation between the magnitude of the generalized order parameters of a given N-H vector and the presence of hydrogen bonding of the amide hydrogen or its peptide bond associated carbonyl. Using a chemical shift tensor breadth of 160 ppm, the N-H vectors of peptide linkages participating in one or more hydrogen bonds to the main chain show an average generalized order parameter of 0.80 (SD 0.06), while those amide NH of peptide linkages free of hydrogen-bonding interactions with the main chain show an average order parameter of 0.69 (SD 0.06).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
The hydrogen-deuterium exchange reactions in gamma-irradiated DL-alanine in the solid state were investigated by spin-trapping and electron spin resonance (e.s.r) using selectively deuterated DL-alanine. Subsequent to gamma-radiolysis at 30 degrees C, polycrystalline DL-alanine was dissolved in aqueous solutions of 2-methyl-2-nitrosopropane and the extent of H-D exchange of the deamination radicals was followed by e.s.r. After formation of the deamination radicals, four exchange reactions were found to occur between the radicals and the surrounding undamaged molecules. The first reaction, which occurs between the hydrogens of the C-2 carbon of the radicals and those of the methyl groups of the neighbouring molecules, can be followed at room temperature. The three other reactions could be conveniently monitored in gamma-irradiated polycrystalline alanine at 110 degrees C. The first of the other three reactions takes place between the methyl hydrogens of the radicals and the C-2 hydrogens of nearby molecules, while the remaining processes involve exchange between the hydrogen atoms of the amino group and those on the C-2 and C-3 carbon atoms of the deamination radical.  相似文献   

10.
Adenosylcobalamin-dependent diol dehydratase of Klebsiella oxytoca is apparently not stereospecific and catalyzes the conversion of both (R)- and (S)-1,2-propanediol to propionaldehyde. To explain this unusual property of the enzyme, we analyzed the crystal structures of diol dehydratase in complexes with cyanocobalamin and (R)- or (S)-1,2-propanediol. (R)- and (S)-isomers are bound in a symmetrical manner, although the hydrogen-bonding interactions between the substrate and the active-site residues are the same. From the position of the adenosyl radical in the modeled "distal" conformation, it is reasonable for the radical to abstract the pro-R and pro-S hydrogens from (R)- and (S)-isomers, respectively. The hydroxyl groups in the substrate radicals would migrates from C(2) to C(1) by a suprafacial shift, resulting in the stereochemical inversion at C(1). This causes 60 degrees clockwise and 70 degrees counterclockwise rotations of the C(1)-C(2) bond of the (R)- and (S)-isomers, respectively, if viewed from K+. A modeling study of 1,1-gem-diol intermediates indicated that new radical center C(2) becomes close to the methyl group of 5'-deoxyadenosine. Thus, the hydrogen back-abstraction (recombination) from 5'-deoxyadenosine by the product radical is structurally feasible. It was also predictable that the substitution of the migrating hydroxyl group by a hydrogen atom from 5'-deoxyadenosine takes place with the inversion of the configuration at C(2) of the substrate. Stereospecific dehydration of the 1,1-gem-diol intermediates can also be rationalized by assuming that Asp-alpha335 and Glu-alpha170 function as base catalysts in the dehydration of the (R)- and (S)-isomers, respectively. The structure-based mechanism and stereochemical courses of the reaction are proposed.  相似文献   

11.
SET domain lysine methyltransferases (KMTs) are S-adenosylmethionine (AdoMet)-dependent enzymes that catalyze the site-specific methylation of lysyl residues in histone and non-histone proteins. Based on crystallographic and cofactor binding studies, carbon-oxygen (CH · · · O) hydrogen bonds have been proposed to coordinate the methyl groups of AdoMet and methyllysine within the SET domain active site. However, the presence of these hydrogen bonds has only been inferred due to the uncertainty of hydrogen atom positions in x-ray crystal structures. To experimentally resolve the positions of the methyl hydrogen atoms, we used NMR (1)H chemical shift coupled with quantum mechanics calculations to examine the interactions of the AdoMet methyl group in the active site of the human KMT SET7/9. Our results indicated that at least two of the three hydrogens in the AdoMet methyl group engage in CH · · · O hydrogen bonding. These findings represent direct, quantitative evidence of CH · · · O hydrogen bond formation in the SET domain active site and suggest a role for these interactions in catalysis. Furthermore, thermodynamic analysis of AdoMet binding indicated that these interactions are important for cofactor binding across SET domain enzymes.  相似文献   

12.
The benefits of performing ENDOR experiments at higher microwave frequency are demonstrated in a Q-band (35 GHz) ENDOR investigation of a number of proteins with [nFe-mS] clusters, n = 2, 3, 4. Each protein displays several resonances in the frequency range of 0-20 MHz. In all instances, features are seen near v approximately 13 and 8 MHz that can be assigned, respectively, to "distant ENDOR" from 13C in natural-abundance (1.1%) and from 14N (the delta m1 = +/- 2 transitions); the nuclei involved in this phenomenon are remote from and have negligible hyperfine couplings to the cluster. In addition, a number of proteins show local 13C ENDOR signals with resolved hyperfine interactions; these are assigned to the beta carbons of cysteines bound to the cluster [A(13C) approximately 1.0 MHz]. Five proteins show resolved, local delta m1 = +/- 2 ENDOR signals from 14N with an isotropic hyperfine coupling, 0.4 less than or equal to A(14N) less than or equal to 1.0, similar to those seen in ESEEM studies; these most likely are associated with N-H...S hydrogen bonds to the cluster. Anabaena ferredoxin further shows a signal corresponding to A(14N) approximately 4 MHz. Quadrupole coupling constants are derived for both local and distant 14N signals. The interpretation of the data is supported by studies on 15N- and 13C-enriched ferredoxin (Fd) from Anabaena 7120, where the 15N signals can be clearly correlated with the corresponding 14N signals and where the 13C signals are strongly enhanced. Thus, the observation of 14N delta m1 = +/- 2 signals at Q-band provides a new technique for examining weak interactions with a cluster.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
The ability of horseradish peroxidase (E.C. 1.11.1.7. Donor: H2O2 oxidoreductase) to catalytically oxidize 2-deoxyribose sugars to a free radical species was investigated. The ESR spin-trapping technique was used to denionstrate that free radical species were formed. Results with the spin trap 3.5-dibronio-4-nitrosoben-zene sulphonic acid showed that horseradish peroxidase can catalyse the oxidation of 2-deoxyribose to produce an ESR spectrum characteristic of a nitroxide radical spectrum. This spectrum was shown to be a composite of spin adducts resulting from two carbon-centered species, one spin adduct being characterized by the hyperfine coupling constants aN = 13.6GandaHβ = 11.0G, and the other by aN = 13.4G and aH = 5.8 G. When 2-deoxyribose-5-phosphate was used as the substrate, the spectrum produced was found to be primarily one species characterized by the hyperfine coupling constants aN = 13.4G and aH= 5.2. All the radical species produced were carbon-centered spin adducts with a β hydrogen, suggesting that oxidation occurred at the C(2) or C(5) moiety of the sugar. Interestingly, it was found that under the same experimental conditions, horseradish peroxidase apparently did not catalyze the oxidation of either 3-deoxyribose or D-ribose to a free radical since no spin adducts were found in these cases.

It can be readily seen that 2-deoxyribose and 2-deoxyribose-5-phosphate can be oxidized by HRP/H2O2 to form a free radical species that can be detected with the ESR spin-trapping technique. There are two probable sites for the formation of a CH type radical on the 2-deoxyribose sugar, these being the C(2) and the C(5) carbons. The fact that there is a species produced from 2-deoxy-ribose, but not 2-deoxy-ribose-5-phosphate, suggests that there is an involvement of the C(5) carbon in the species with the 1 1.0G β hydrogen. In the spectra formed from 2-deoxy-ribose, there is a big difference in the hyperfine splitting of the β hydrogens, suggesting that the radicals are formed at different carbon centers, while the addition of a phosphate group to the C(5) carbon seems to inhibit radical formation at one site. In related work, the chemiluminescence of monosaccharides in the presence of horseradish peroxidase was proposed to be the consequence of carbon-centered free radical formation (10).  相似文献   

14.
The conformation of the N-linked complex glycopeptide of fetuin was examined with hydrogen-exchange techniques. The glycopeptide molecule contains eight acetamido hydrogens stemming from five N-acetylglucosamine residues and three N-acetylneuraminic acid residues and also one from the remaining sugar-peptide linkage. The hydrogen-exchange rates of these secondary amides were compared with small molecule model compounds having identical primary structures at their exchangeable hydrogen sites. Differences between the model rates and glycopeptide rates therefore cannot be accounted for by primary structure effects but reflect conformational features of the glycopeptide. Two glycopeptide hydrogens exhibit significantly hindered exchange; the rest exchange at the model rates. Removal of the three N-acetylneuraminic acid residues from terminal positions on the three branches of the glycopeptide removes the slowed hydrogens. The remaining ones continue to exchange at the model rate. These results indicate that two of the eight sugar acetamido hydrogens are involved in intramolecular hydrogen bonds. A likely structure includes two hydrogen bonds between the three N-acetylneuraminic acid residues. These two hydrogens, slowed to a moderate degree, reflect a preferred conformation stabilized by about 1 kcal/mol in free energy. The solution conformation of the glycopeptide suggested by these results is one that is partially ordered and can be easily modulated, owing to the relatively small amount of energy stabilizing the preferred conformation.  相似文献   

15.
The reaction of tryptamine with indolyl-3-alkane alpha-hydroxylase is shown to remove stereospecifically the pro-S hydrogen at C-2 of the side chain and to give hydroxytryptamine of "R" configuration. The reaction therefore proceeds stereospecifically with net inversion of configuration at C-2 of the tryptamine side chain. In the reaction of L-tryptophan methyl ester, the enzyme also catalyzes stereospecific removal of the pro-S hydrogen at C-3, but the product 3-hydroxytryptophan methyl ester is racemic at C-3. The unreacted tryptophan methyl ester is shown to incorporate solvent hydrogen into the pro-S position at C-3 in an at least partially stereospecific manner, suggesting that the reaction of L-tryptophan methyl ester is reversible. The hydrogens at C-1 of the tryptamine side chain and the alpha-hydrogen of L-tryptophan methyl ester are shown to be retained in the reactions. The results support the notion that the enzyme catalyzes stereospecific 1,4-dehydrogenation of 3-substituted indoles to the coresponding alkylidene indolenines as the primary reaction, followed by stereospecific or nonstereospecific hydration of these intermediates as a secondary process. Substrate specificity studies with a number of tryptophan analogs are in excellent agreement with such a mechanism.  相似文献   

16.
Isothermal dielectric loss spectra of neutralized and nonneutralized chitosan were acquired in successive runs from -130 degrees C up to increasing final temperatures, in a frequency range between 20 Hz and 1 MHz. Essentially, three relaxation processes were detected in the temperature range covered: (i) a beta-wet process, detected when the sample has a higher water content that vanishes after heating to 150 degrees C; (ii) a beta process, which is located at temperatures below 0 degrees C, becoming better defined and maintaining its location after annealing at 150 degrees C independently of the protonation state of the amino side group; and (iii) a sigma process that deviates to higher temperatures with drying, being more mobile in the nonneutralized form. Moreover, in dried neutralized chitosan, a fourth process was detected in the low frequency side of the secondary beta process that diminishes after annealing. Whether this process is a distinct relaxation of the dried polymer or a deviated beta-wet process due to the loss of water residues achieved by annealing is not straightforward. Only beta and sigma processes persist after annealing at 150 degrees C. The changes in molecular mobility upon drying of these two relaxation processes were evaluated.  相似文献   

17.
《Free radical research》2013,47(3-6):297-302
The ability of horseradish peroxidase (E.C. 1.11.1.7. Donor: H2O2 oxidoreductase) to catalytically oxidize 2-deoxyribose sugars to a free radical species was investigated. The ESR spin-trapping technique was used to denionstrate that free radical species were formed. Results with the spin trap 3.5-dibronio-4-nitrosoben-zene sulphonic acid showed that horseradish peroxidase can catalyse the oxidation of 2-deoxyribose to produce an ESR spectrum characteristic of a nitroxide radical spectrum. This spectrum was shown to be a composite of spin adducts resulting from two carbon-centered species, one spin adduct being characterized by the hyperfine coupling constants aN = 13.6GandaHβ = 11.0G, and the other by aN = 13.4G and aH = 5.8 G. When 2-deoxyribose-5-phosphate was used as the substrate, the spectrum produced was found to be primarily one species characterized by the hyperfine coupling constants aN = 13.4G and aH= 5.2. All the radical species produced were carbon-centered spin adducts with a β hydrogen, suggesting that oxidation occurred at the C(2) or C(5) moiety of the sugar. Interestingly, it was found that under the same experimental conditions, horseradish peroxidase apparently did not catalyze the oxidation of either 3-deoxyribose or D-ribose to a free radical since no spin adducts were found in these cases.

It can be readily seen that 2-deoxyribose and 2-deoxyribose-5-phosphate can be oxidized by HRP/H2O2 to form a free radical species that can be detected with the ESR spin-trapping technique. There are two probable sites for the formation of a CH type radical on the 2-deoxyribose sugar, these being the C(2) and the C(5) carbons. The fact that there is a species produced from 2-deoxy-ribose, but not 2-deoxy-ribose-5-phosphate, suggests that there is an involvement of the C(5) carbon in the species with the 1 1.0G β hydrogen. In the spectra formed from 2-deoxy-ribose, there is a big difference in the hyperfine splitting of the β hydrogens, suggesting that the radicals are formed at different carbon centers, while the addition of a phosphate group to the C(5) carbon seems to inhibit radical formation at one site. In related work, the chemiluminescence of monosaccharides in the presence of horseradish peroxidase was proposed to be the consequence of carbon-centered free radical formation (10).  相似文献   

18.
Side chain dynamics monitored by 13C-13C cross-relaxation   总被引:1,自引:0,他引:1  
A method to measure (13)C-(13)C cross-relaxation rates in a fully (13)C labeled protein has been developed that can give information about the mobility of side chains in proteins. The method makes use of the (H)CCH-NOESY pulse sequence and includes a suppression scheme for zero-quantum (ZQ) coherences that allows the extraction of initial rates from NOE buildup curves.The method has been used to measure (13)C-(13)C cross-relaxation rates in the 269-residue serine-protease PB92. We focused on C(alpha)-C(beta) cross-relaxation rates, which could be extracted for 64% of all residues, discarding serine residues because of imperfect ZQ suppression, and methyl (13)C-(13)C cross-relaxation rates, which could be extracted for 47% of the methyl containing C-C pairs. The C(alpha)-C(beta) cross-relaxation rates are on average larger in secondary structure elements as compared to loop regions, in agreement with the expected higher rigidity in these elements. The cross-relaxation rates for methyl containing C-C pairs show a general decrease of rates further into the side chain, indicating more flexibility with increasing separation from the main chain. In the case of leucine residues also long-range C(beta)-C(delta) cross-peaks are observed. Surprisingly, for most of the leucines a cross-peak with only one of the methyl C(delta) carbons is observed, which correlates well with the chi(2) torsion-angle and can be explained by a difference in mobility for the two methyl groups due to an anisotropic side chain motion.  相似文献   

19.
ACTH (1-10), an adrenocorticotropin hormone fragment, was studied by molecular dynamics (MD) simulation in the NPT ensemble in an explicit sodium dodecylsulfate (SDS) micelle. Initially, distance restraints derived from NMR nuclear Overhauser enhancements were incorporated during the equilibration stage of the simulation. The analyses of the trajectories from the subsequent unrestrained MD showed that ACTH (1-10) does not conform to a helical structure at the micelle-water interface; however, the structure is amphipathic. The loss of the helical structure is due to decreased intramolecular hydrogen bonding accompanied by an increase of hydrogen bonding between the amide hydrogens of the peptide and the micelle head-groups. ACTH (1-10) was found to lie on the surface of the SDS micelle. Most of the hydrophobic interactions came from the side-chains of Met-4, Phe-7 and Trp-9. The peptide bonds were either hydrated or involved in intramolecular hydrogen bonding. Decreased hydration for the backbone of His-6 and Phe-7 was due to intermolecular hydrogen bonding with the SDS head-groups. The time correlation functions of the N-H bonds of the peptide in water and in the micelle showed that the motions of the peptide, except for the N- and C-termini, are significantly reduced when partitioned in the micelle.  相似文献   

20.
Redox state-dependent changes in the relative orientation of the phenol side chain and the peptide group in model tyrosine have been characterized using specific 2H isotopic labelling and X-band electron paramagnetic resonance (EPR) spectroscopy. Tyrosyl radicals were generated by UV photolysis of tyrosine trapped in rigid polycrystalline basic-aqueous medium at T < or = 170 K. Ring-2H(4) and beta-2H(2) substitutions on tyrosine were used to enhance the lineshape contributions from beta-hydrogen or ring-hydrogen hyperfine interactions, respectively. The EPR lineshape at 120 K of the trapped ring-2H(4)-tyrosyl radical is altered dramatically after annealing at 235 K. In contrast, the lineshape of the beta-2H(2)-tyrosyl radical is impervious to annealing. The effect of annealing on the lineshape therefore arises from a change in the isotropic hyperfine coupling between unpaired pi-electron spin density at the ring carbon atom C(1) and the beta-hydrogen nuclei, which is caused by rotational relaxation of the ring and peptide group about the C(1)-C(beta) bond. EPR simulations indicate angular distributions of the peptide group (R-) of 0 degrees < or = theta(R) < or = 30 degrees and 0 degrees < or = theta(R)< or = 18 degrees in the rigid and relaxed radical states, respectively. Redox-induced changes in the C(1)-C(beta) rotamer distribution must be accounted for in assessments of stable amino acid side chain equilibrium structures, and may influence catalytic tyrosyl radical/tyrosine function in enzymes.  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