首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Proteinaceous residues incorporated within the crystal structure of ostrich eggshells (OES) are retained without loss over geological time exceeding 10 million years. Degradation of the polypeptides, including hydrolysis to smaller peptide fragments and eventual release of free amino acids, decomposition, and racemization and epimerization occur at regular, predictable rates dependent on ambient temperature. The extent of isoleucine epimerization (aIle/Ile ratio) in OES follows linear first-order reversible kinetics in controlled-temperature laboratory simulations of time up to an aIle/Ile ratio in excess of 1.0. The hydrolysis of leucine also follows a predictable pattern, but deviates from first-order kinetics. A nonlinear mathematical model has been developed that adequately describes the pattern of leucine hydrolysis through a wide temperature range. Arrhenius parameters were derived from laboratory experiments combined with rate constant values found for 14C-dated OES from stratified caves in southern Africa. These parameters for isoleucine epimerization and leucine hydrolysis differ by ca. 10%, allowing the simultaneous solution of the two equations for temperature, independent of sample age. Although the uncertainty of the simultaneous temperature is relatively high (+/- 10 degrees C), it provides an effective means of identifying burned samples. If sample age is known, palaeotemperatures (the integrated thermal history experienced by an eggshell as opposed to an 'instantaneous' temperature) can be calculated with a precision of better than +/- 1 degrees C. The ages of levels at Border Cave, South Africa, from which anatomically modern human skeletal remains have been recovered, are dated by the extent of isoleucine epimerization in associated OES.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
BACKGROUND: Methylmalonyl-CoA epimerase (MMCE) is an essential enzyme in the breakdown of odd-numbered fatty acids and of the amino acids valine, isoleucine, and methionine. Present in many bacteria and in animals, it catalyzes the conversion of (2R)-methylmalonyl-CoA to (2S)-methylmalonyl-CoA, the substrate for the B12-dependent enzyme, methylmalonyl-CoA mutase. Defects in this pathway can result in severe acidosis and cause damage to the central nervous system in humans. RESULTS: The crystal structure of MMCE from Propionibacterium shermanii has been determined at 2.0 A resolution. The MMCE monomer is folded into two tandem betaalphabetabetabeta modules that pack edge-to-edge to generate an 8-stranded beta sheet. Two monomers then pack back-to-back to create a tightly associated dimer. In each monomer, the beta sheet curves around to create a deep cleft, in the floor of which His12, Gln65, His91, and Glu141 provide a binding site for a divalent metal ion, as shown by the binding of Co2+. Modeling 2-methylmalonate into the active site identifies two glutamate residues as the likely essential bases for the epimerization reaction. CONCLUSIONS: The betaalphabetabetabeta modules of MMCE correspond with those found in several other proteins, including bleomycin resistance protein, glyoxalase I, and a family of extradiol dioxygenases. Differences in connectivity are consistent with the evolution of these very different proteins from a common precursor by mechanisms of gene duplication and domain swapping. The metal binding residues also align precisely, and striking structural similarities between MMCE and glyoxalase I suggest common mechanisms in their respective epimerization and isomerization reactions.  相似文献   

3.
Peptide aldehydes are interesting targets as enzyme inhibitors, and can be used for pseudopeptide chemistry or ligation. However, they are known to be subjected to epimerization during synthesis or purification. By (1)H NMR, a model dipeptide aldehyde can be used to check the possible epimerization occurring during synthesis. Various purification methods were investigated, but none was free from epimerization.  相似文献   

4.
Quantitative estimates of time-averaging (age mixing) in gastropod shell accumulations from Quaternary (the late Pleistocene and Holocene) eolian deposits of Canary Islands were obtained by direct dating of individual gastropods obtained from exceptionally well-preserved dune and paleosol shell assemblages. A total of 203 shells of the gastropods Theba geminata and T. arinagae, representing 44 samples (= stratigraphic horizons) from 14 sections, were dated using amino acid (isoleucine) epimerization ratios calibrated with 12 radiocarbon dates. Most samples reveal a substantial variation in shell age that exceeds the error that could be generated by dating imprecision, with the mean within-sample shell age range of 6670 years and the mean standard deviation of 2920 years. Even the most conservative approach (Monte Carlo simulations with a non-sequential Bonferroni correction) indicates that at least 25% of samples must have undergone substantial time-averaging (e.g., age variations within those samples cannot be explained by dating imprecision alone). Samples vary in shell age structure, including both left-skewed (17 out of 44) and right-skewed distributions (26 out of 44) as well as age distributions with a highly variable kurtosis. Dispersion and shape of age distributions of samples do not show any notable correlation with the stratigraphic age of samples, suggesting that the structure and scale of temporal mixing is time invariant. The statistically significant multi-millennial time-averaging observed here is consistent with previous studies of shell accumulations from various depositional settings and reinforces the importance of dating numerous specimens per horizon in geochronological studies. Unlike in the case of marine samples, typified by right-skewed age distributions (attributed to an exponential-like shell loss from older age classes), many of the samples analyzed here displayed left-skewed distributions, suggestive of different dynamics of age mixing in marine versus terrestrial shell accumulations.  相似文献   

5.
Guo C  Shah RD  Mills J  Dukor RK  Cao X  Freedman TB  Nafie LA 《Chirality》2006,18(10):775-782
Near-infrared (near-IR) Fourier transform vibrational circular dichroism (FT-VCD) spectroscopy has been used to monitor the epimerization of (S)-(+)-2,2-dimethyl-1,3-dioxolane-4-methanol (S-DDM). The near-IR-VCD spectra display clear isolated VCD bands at the range of 4700-5050 cm(-1) resulting from the OH stretch-bend combination bands of S-DDM, which were found to decrease in intensity with increasing reaction time. The near-IR-VCD spectra of 10 reference samples obtained were subjected to partial least-squares (PLS) regression analysis, and the results were used to build predictive models for enantiomeric excess (EE) determination. Multivariate regression was carried out on three different sets of spectra, corresponding to the epimerization of S-DDM in three different solvents: methylcyclohexane, carbon tetrachloride and tetrahydrofuran. The effects of solvent in DDM epimerization are discussed in terms of the relative stabilization of the reaction intermediate of the DDM epimerization reaction. The results of these near-IR-VCD studies for the determination of EE highlights the potential of VCD for in situ real-time process monitoring of the reaction kinetics of chiral molecules in solution.  相似文献   

6.
Abstract The fermentation of isoleucine, arginine and isoleucine + arginine by pure and syntrophic cultures of Clostridium sporogenes was investigated. Growth of C. sporogenes on isoleucine, if any, was poor, but some isoleucine was fermented to 2-methylbutyrate and hydrogen. In syntrophic cultures with Methanobacterium formicicum or Methanosarcina barkeri growth was better, and isoleucine was completely fermented, the hydrogen being used for methane production. Pure cultures of C. sporogenes grew on arginine and produced 5-aminovalerate, ornithine and acetate. The reducing equivalents for 5-aminovalerate production from intermediarily formed proline were provided by oxidative conversion of arginine to acetate and by oxidative metabolism of some amino acids present in the yeast extract. However, when isoleucine was available together with arginine in syntrophic cultures of C. sporogenes and M. formicicum , the reducing equivalents for arginine fermentation came mainly from the oxidation of isoleucine (Stickland reaction), and the hydrogen produced in excess served for the reduction of CO2 to methane.  相似文献   

7.
Recently, epimerization of the hydroxyl group at C-3 has been identified as a unique metabolic pathway of vitamin D compounds. We measured C-3 epimerization activity in subcellular fractions prepared from cultured cells and investigated the basic properties of the enzyme responsible for the epimerization. C-3 epimerization activity was detected using a NADPH-generating system containing glucose-6-phosphate, NADP, glucose-6-phosphate dehydrogenase, and Mg(2+). The highest level of activity was observed in a microsomal fraction prepared from rat osteoblastic UMR-106 cells but activity was also observed in microsomal fractions prepared from MG-63, Caco-2, Hep G2, and HUH-7 cells. In terms of maximum velocity (V(max)) and the Michaelis constant (K(m)), 25-hydroxyvitamin D(3) [25(OH)D(3)] exhibited the highest specificity for the epimerization at C-3 among 1alpha,25-dihydroxyvitamin D(3) [1alpha,25(OH)(2)D(3)], 25(OH)D(3), 24,25-dihydroxyvitamin D(3) [24,25(OH)(2)D(3)], and 22-oxacalcitriol (OCT). The epimerization activity was not inhibited by various cytochrome P450 inhibitors and antiserum against NADPH cytochrome P450 reductase. Neither CYP24, CYP27A1, CYP27B1 nor 3(alpha-->beta)hydroxysteroid epimerase (HSE) catalyzed the epimerization in vitro. Based on these results, the enzyme(s) responsible for the epimerization of vitamin D(3) at C-3 are thought to be located in microsomes and different from cytochrome P450 and HSE.  相似文献   

8.
We investigated the mechanism of epimerization (R to S or S to R) of moxalactam in serum of rats, dogs, and humans. The epimerization of moxalactam occurred in the serum of these animals, but not in the serum filtrate. The albumin fraction of human serum purified by gel filtration catalysed the epimerization of moxalactam at an identical rate to serum, but other fractions (i.e., lipoproteins and globulins) showed slower epimerization. alpha 1-acid glycoprotein, which was eluted in the same fraction with albumin by G-200 gel filtration, did not epimerize moxalactam. The presence of 2 mM warfarin decreased the binding of R- and S-moxalactam and decreased the epimerization of moxalactam in human serum. These results demonstrate moxalactam was epimerized on the warfarin binding site on albumin in serum. Additionally, a physiologically based pharmacokinetic model shows that the epimerization of moxalactam after administration in dogs is simulated by the epimerization in serum.  相似文献   

9.
Paracoccus denitrificans was grown on either [2,3-13C]succinate or [1,4-13C]succinate, and extracts were analysed by using gas chromatography-mass spectrometry. The distribution of label in isoleucine indicated that the 2-ketobutyrate required for isoleucine biosynthesis was mainly produced from pyruvate by 2-keto-acid chain elongation (i.e. the 'pyruvate elongation pathway'). Approximately 10% of isoleucine was produced by a second pathway involving propionyl CoA. Threonine and glutamate were not utilized by P. denitrificans as a source of 2-ketobutyrate production for isoleucine biosynthesis under the growth conditions used.  相似文献   

10.
Chen D  Abend A  Stubbe J  Frey PA 《Biochemistry》2003,42(15):4578-4584
The adenosylcobalamin-dependent ribonucleoside triphosphate reductase (RTPR) from Lactobacillus leichmannii catalyzes the reduction of ribonucleoside triphosphates to deoxyribonucleoside triphosphates. RTPR also catalyzes the exchange of the C5'-hydrogens of adenosylcobalalamin with solvent hydrogen. A thiyl radical located on Cys 408 is generated by reaction of adenosylcobalamin at the active site and is proposed to be the intermediate for both the nucleotide reduction and the 5'-hydrogen exchange reactions. In the present research, a stereochemical approach is used to study the mechanism of the Co-C5' bond cleavage of adenosylcobalamin in the reaction of RTPR. When stereoselectively deuterated coenzyme, (5'R)-[5'-(2)H(1)] adenosylcobalamin (5'R/S = 3:1), was incubated with RTPR or the Cys 408 viariants, C408A-RTPR and C408S-RTPR in the presence of dGTP, the deuterium at the 5'-carbon was stereochemically scrambled, leading to epimerization of the (5'S)-[5'-(2)H(1)]- and (5'R)-[5'-(2)H(1)]-isotopomers. Observation of epimerization with mutated RTPR proves that transient cleavage of the Co-C5' bond occurs in the absence of the thiol group on Cys 408. The rate constants for epimerization by RTPR, C408A-RTPR, and C408S-RTPRs in the presence of dGTP are 5.1, 0.28, and 0.42 s(-1), respectively. Only the wild-type RTPR catalyzes the 5'-hydrogen exchange reaction. Both epimerization and 5'-hydrogen exchange reactions are stimulated by the allosteric effector dGTP, and epimerization is not detected in the absence of the effector. Mechanistic implications with respect to wt-RTPR-mediated carbon cobalt bond homolysis and the intermediacy of the 5'-deoxyadenosyl radical will be presented.  相似文献   

11.
Wang Y  Li Y  Wu Y  Yan H 《The FEBS journal》2007,274(9):2240-2252
Dihydroneopterin aldolase (DHNA) catalyzes both the cleavage of 7,8-dihydro-D-neopterin (DHNP) to form 6-hydroxymethyl-7,8-dihydropterin (HP) and glycolaldehyde and the epimerization of DHNP to form 7,8-dihydro-L-monapterin (DHMP). Whether the epimerization reaction uses the same reaction intermediate as the aldol reaction or the deprotonation and reprotonation of C2' of DHNP has been investigated by NMR analysis of the reaction products in a D2O solvent. No deuteration of C2' was observed for the newly formed DHMP. This result strongly suggests that the epimerization reaction uses the same reaction intermediate as the aldol reaction. In contrast with an earlier observation, the DHNA-catalyzed reaction is reversible, which also supports a nonstereospecific retroaldol/aldol mechanism for the epimerization reaction. The binding and catalytic properties of DHNAs from both Staphylococcus aureus (SaDHNA) and Escherichia coli (EcDHNA) were determined by equilibrium binding and transient kinetic studies. A complete set of kinetic constants for both the aldol and epimerization reactions according to a unified kinetic mechanism was determined for both SaDHNA and EcDHNA. The results show that the two enzymes have significantly different binding and catalytic properties, in accordance with the significant sequence differences between them.  相似文献   

12.
13.
Jerga A  Stanley MD  Tipton PA 《Biochemistry》2006,45(30):9138-9144
C5-mannuronan epimerase catalyzes the formation of alpha-L-guluronate residues from beta-D-mannuronate residues in the synthesis of the linear polysaccharide alginate. The reaction requires the abstraction of a proton from C5 of the residue undergoing epimerization followed by re-protonation on the opposite face. Rapid-mixing chemical quench experiments were conducted to determine the nature of the intermediate formed upon proton abstraction in the reaction catalyzed by the enzyme from Pseudomonas aeruginosa. Colorimetric and HPLC analysis of quenched samples indicated that shortened oligosaccharides containing an unsaturated sugar residue form as transient intermediates in the epimerization reaction. This suggests that the carbanion is stabilized by glycal formation, concomitant with cleavage of the glycosidic bond between the residue undergoing epimerization and the adjacent residue. The time dependence of glycal formation suggested that slow steps flank the chemical steps in the catalytic cycle. Solvent isotope effects on V and V/K were unity, consistent with a catalytic cycle in which chemistry is not rate-limiting. The specificity of the epimerase with regard to neighboring residues was examined, and it was determined that the enzyme showed no bias for mannuronate residues adjacent to guluronates versus those adjacent to mannuronates. Proton abstraction and sugar epimerization were irreversible. Existing guluronate residues already present in the polysaccharide were not converted to mannuronates, nor was incorporation of solvent deuterium into existing mannuronates observed.  相似文献   

14.
Interconversion of D- and L-isomers of 3-hydroxy-decanoyl-CoA was catalyzed by rat liver homogenate. Cation exchange chromatography followed by ammonium sulfate precipitation and PBE-94 chromatofocusing column was used to separate the peroxisomal bifunctional protein, the classic 2-enoyl-CoA hydratase (crotonase), and a novel 2-enoyl-CoA hydratase. Epimerization activity was lost during the last purification step. None of the above proteins was capable of catalyzing the epimerization by itself, but reconstitution was achieved by recombining crotonase and the novel 2-enoyl-CoA hydratase. Since hydration by the latter enzyme follows a different stereochemical course from that with crotonase, these two hydratases are distinguished as 2-enoyl-CoA hydratase 1 (crotonase) and 2-enoyl-CoA hydratase 2 (the novel hydratase). The data strongly suggested that epimerization in the rat liver proceeds via dehydration-hydration catalyzed by the two different hydratases. The intermediate of this two step mechanism appears to be trans-2-enoyl-CoA.  相似文献   

15.
An epimerization-free system for coupling N-protected peptides with free amino acids was developed. A number of inorganic substances were tested as epimerization suppressant additives during the coupling by various methods (carbodiimide plus additives, uronium salts, Woodward's reagent-K, isobutyl-chloroformate, etc.). Some of them (ZnCl2, RbClO4, LiCl, SnCl4, AlCl3, etc.) in combination with some coupling methods can guarantee coupling with minimal epimerization (D-epimer < 1%). But only a simultaneous use of 1-hydroxybenzotriazole and Cu2+ ions as additives in carbodiimide-mediated peptide couplings appeared to give a standard result (D-epimer < 0.1%). There was no epimerization even in the case when N-methyl amino acid (sarcosine) was used as an amino component, while in the absence of Cu2+ ions an unacceptable level of epimerization was observed (D-epimer, 22% for carbodiimide with the 1-hydroxybenzotriazole method). So far it has been considered that Cu2+ ions prevent obtaining peptides in high yields (< 90%) by various coupling methods. We have found that the use of 1-hydroxybenzotriazole, CuCl2 and 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide instead of dicyclohexylcarbodiimide provides a possible method for obtaining the desired peptides in 90-99% yields without epimerization. All these results were shown by employing several model peptide couplings with free amino acids as amino components dissolved in an effective solvent system which readily dissolved them.  相似文献   

16.
SAR studies on the stereoisomers of CP-470,711 suggested that in vivo epimerization was taking place in rats. Further metabolism studies revealed that no epimerization was occurring in dogs, and that no epimerization was expected in humans. A mechanism for the in vivo epimerization is proposed involving an oxidation-reduction pathway of the secondary benzylic alcohol, in contrast to an acid/base-promoted epimerization of the same center during chemical synthesis.  相似文献   

17.
Galactose depleted locust bean gum was selectively oxidized in C(6) position and epimerized with mannuronan C(5)-epimerases to obtain the corresponding artificial uronanes. These new pseudo-alginates were characterized by NMR spectroscopy and circular dichroism (CD). Specifically, 1D and 2D NMR techniques allowed the degree of epimerization, the distribution of mannuronic and guluronic acid residues in the polysaccharidic chain, and the average G block length to be determined. In addition, NMR diffusion experiments showed that the epimerization reaction did not significantly degrade the polysaccharidic chains. Circular dichroism was used to investigate the kinetics of the epimerization reaction and to evidence the specific interaction between the epimerized locust bean samples with Ca(II) ions in dilute solution. All of the samples considered in this study form wall to wall gels in concentrated polymer solutions.  相似文献   

18.
Y H Lim  K Yokoigawa  N Esaki    K Soda 《Journal of bacteriology》1993,175(13):4213-4217
We have found that Pseudomonas putida ATCC 17642 cells grown in a medium containing D-threonine as the sole nitrogen source produce an enzyme that catalyzes epimerization of threonine. Proton nuclear magnetic resonance analysis of the enzyme reaction in deuterium oxide clearly showed epimerization from L- to D-allo-threonine and also from D- to L-allo-threonine. This is the first example of an enzyme that was clearly shown to epimerize threonine. The enzyme has been purified to homogeneity, which was shown by polyacrylamide gel electrophoresis. The enzyme has a molecular weight of about 82,000 and consists of two subunits identical in molecular weight (about 41,000). The enzyme contains 1 mol of pyridoxal 5'-phosphate per mol of subunit as a cofactor, and its absorption spectrum exhibits absorption maxima at 280 and 420 nm. The enzyme catalyzes not only epimerization of threonine by stereoconversion at the alpha position but also racemization of various amino acids, except acidic and aromatic amino acids. The enzyme is similar to amino acid racemase with low substrate specificity (EC 5.1.1.10) in enzymological properties but is distinct from it in the action on threonine.  相似文献   

19.
Peptide segment couplings are now widely utilized in protein chemical synthesis. One of the key structures for the strategy is the peptide thioester. Peptide thioester condensation, in which a C‐terminal peptide thioester is selectively activated by silver ions then condensed with an amino component, is a powerful tool. But the amino acid adjacent to the thioester is at risk of epimerization. During the preparation of peptide thioesters by the Boc solid‐phase method, no substantial epimerization of the C‐terminal amino acid was detected. Epimerization was, however, observed during a thioester–thiol exchange reaction and segment condensation in DMSO in the presence of a base. In contrast, thioester–thiol exchange reactions in aqueous solutions gave no epimerization. The epimerization during segment condensation was significantly suppressed with a less polar solvent that is applicable to segments in thioester peptide condensation. These results were applied to a longer peptide thioester condensation. The epimer content of the coupling product of 89 residues was reduced from 27% to 6% in a condensation between segments of 45 and 44 residues for the thioester and the amino component, respectively. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
J N Batist  A F Slobbe  A F Marx 《Steroids》1989,54(3):321-332
Practical routes to 9 alpha-hydroxypregnenes were developed by epimerization and hydration of 17 alpha-ethynyl-9 alpha,17 beta-dihydroxyandrost-4-en-3-one. In the three different methods of epimerization which were used, the C-9 alpha hydroxy group was not susceptible to rearrangement or other side reactions. C-21 functionalized 9 alpha-hydroxypregnenes were obtained by introducing a 17 alpha-halogenated ethynyl group into 9 alpha-hydroxyandrost-4-ene-3,17-dione. Epimerization and hydration by the 17 beta-nitrooxy method produced 21-halogenated 9 alpha-hydroxypregnenes, which were further converted into 21-acetoxy-9 alpha-hydroxypregn-4-ene-3,20-dione.  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