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.
Aminoacyl-tRNA synthetases (AARSs) are ligases (EC.6.1.1.-) that catalyze the acylation of amino acids to their cognate tRNAs in the process of translating genetic information from mRNA to protein. Their amino acid and tRNA specificity are crucial for correctly translating the genetic code. Glycine is the smallest amino acid and the glycyl-tRNA synthetase (GlyRS) belongs to Class II AARSs. The enzyme is unusual because it can assume different quaternary structures. In eukaryotes, archaebacteria and some bacteria, it forms an ??2 homodimer. In some bacteria, GlyRS is an ??2??2 heterotetramer and shows a distant similarity to ??2 GlyRSs. The human pathogen eubacterium Campylobacter jejuni GlyRS (CjGlyRS) is an ??2??2 heterotetramer and is similar to Escherichia coli GlyRS; both are members of Class IIc AARSs. The two-step aminoacylation reaction of tetrameric GlyRSs requires the involvement of both ??- and ??-subunits. At present, the structure of the GlyRS ??2??2 class and the details of the enzymatic mechanism of this enzyme remain unknown. Here we report the crystal structures of the catalytic ??-subunit of CjGlyRS and its complexes with ATP, and ATP and glycine. These structures provide detailed information on substrate binding and show evidence for a proposed mechanism for amino acid activation and the formation of the glycyl-adenylate intermediate for Class II AARSs.  相似文献   

2.
The transglucosylation reaction of buckwheat α-glucosidase was examined under the coexistence of 2-deoxy-d-glucose and maltose. As the transglucosylation products, two kinds of new disaccharide were chromatographically isolated in a crystalline form (hemihydrate). It was confirmed that these disaccharides were 3-O-α-d-glucopyranosyl-2-deoxy-d-glucose ([α]d + 132°, mp 130 ~ 132°C, mp of ±-heptaacetate 151 ~ 152°C) and 4-O-±-d-glucopyranosyl-2-deoxy-d-glucose ([±]d + 136°, mp 168 ~ 170°C), respectively. The principal product formed in the enzyme reaction was 3-O-±-d-glucopyranosyl-2-deoxy-d-glucose.  相似文献   

3.
Summary Obvious protection of the catalytic activity of Esch. coli L-asparaginase by 2-macroglobulin (2M) was observed under conditions otherwise propitious to the dissociation of the tetrameric molecule into inactive subunits, i.e. very diluted enzyme solutions or the presence of either SDS or urea. The degree of protection depended on enzyme and 2M concentrations respectively, and on the preincubation time of the 2M-enzyme mixture prior to substrate addition. The formation of a catalytically active complex between 2M and L-asparaginase was confirmed by gel filtration on a Sephadex-G column and by polyacrylamide gel electrophoresis. The fact that the migration distance of the active complex corresponded to the migration of 2M and the absence in that case of a migration band corresponding to the intact molecule suggest that complexing of the enzyme with 2M prevented its dissociation into subunits and thus its inactivation. Addition of 2M to the already dissociated enzyme molecule did not restore its catalytic activity.Alpha2-macroglobulin was shown to have an inhibiting effect on the proteolytic activity of almost all proteases and no effect on their esterolytic activity. Furthermore, it prevents the inhibition of esterolytic activity by some natural compounds1–5. The effect of 2M on other types of catalytic activity has not been investigated enough to afford a generalization of the possible role of this macroglobulin in the control of enzyme activity in the body.This paper reports the results of an in vitro study of the effect of 2M on the catalytic activity of an important amidase, i.e. L-asparaginase (L-asparagine amidohydrolase 3.5.1.1), which in recent years has been used in the treatment of acute lymphocytic leukemia in children6,7.Abbreviations 2M 2-macroglobulin - E enzyme - SDS sodium dodecylsulfate Part of the results were reported at the 10th International Congress of Biochemistry, Hamburg 1976, Abst. p. 377.  相似文献   

4.
The molecular weights calculated from the amino acid sequences of the A and B chains of the lens protein -crystallin differ only slightly (19830 and 20070, respectively). SDS gel electrophoresis of these chains and comparison with marker proteins yield apparent molecular weights of 19500 for A and 22500 for B. The discrepancy between the value of 22500 and the real molecular weight of 20070 for B vanishes by the combined use of SDS and 6 M urea in the polyacrylamide gels.  相似文献   

5.
A model is presented that describes all the saccharides that are produced during the hydrolysis of starch by an -amylase. Potato amylopectin, the substrate of the hydrolysis reaction, was modeled in a computer matrix. The four different subsite maps presented in literature for -amylase originating from Bacillus amyloliquefaciens were used to describe the hydrolysis reaction in a Monte Carlo simulation. The saccharide composition predicted by the model was evaluated with experimental values. Overall, the model predictions were acceptable, but no single subsite map gave the best predictions for all saccharides produced. The influence of an (16) linkage on the rate of hydrolysis of nearby (14) linkages by the -amylase was evaluated using various inhibition constants. For all the subsites considered the use of inhibition constants led to an improvement in the predictions (a decrease of residual sum of squares), indicating the validity of inhibition constants as such. As without inhibition constants, no single subsite map gave the best fit for all saccharides. The possibility of generating a hypothetical subsite map by fitting was therefore investigated. With a genetic algorithm it was possible to construct hypothetical subsite maps (with inhibition constants) that gave further improvements in the average prediction for all saccharides. The advantage of this type of modeling over a regular fit is the additional information about all the saccharides produced during hydrolysis, including the ones that are difficult to measure experimentally.  相似文献   

6.
An alkali-pretreated gelatin (pI~4.9) was fractionated by means of alcohol coacervation and semi-preparative gel chromatography. The thermal responses of the isolated α fractions, the coacervate and the total gelatin were investigated by 2D-correlation FTIR spectroscopy in the amide I band region (1600-1700 cm?1). The gelation temperature was the same for all examined samples (24.5°C) while the melting temperature of the α? fraction was lower (30°C) than that of the other samples (32.5°C). The 2D COS plots indicate that on cooling (gelation) the core sequence of conformational changes is the same for all samples. On heating, however, the α? fraction deviates from the α?-containing samples and shows an earlier disappearance of the triple helix signal in the event sequence. The lower melting temperature (less thermostable gelatin gel) of the α? fraction thus results from a different conformational cascade of the α? chains upon melting. In all samples the initial conformational changes take place in the β-turns, providing further evidence for the models proposed previously.  相似文献   

7.
Summary We have used the powerful methodology of DNA enzymatic amplification in order to assign human -globin structural mutants to one of the two highly homologous -globin genes. Selectively amplified 1 and 2-globin cDNAs were dot-blotted and further hybridized to synthetic oligonucleotides encompassing either the normal or the mutated sequences. The generated signals corresponded specifically to one of the two -globin genes. Using this approach the -globin structural mutants J-Buda and G-Pest were found to be encoded by the 2 and the 1-globin genes, respectively. Furthermore, the exact nucleotide changes were determined. We propose this technique to serve as a simple and definitive method for assigning -globin structural mutants.  相似文献   

8.
9.
The regions of the HLA-A2 molecule controlling anti-A2 alloreactivity were explored using naturally occurring allelic variants of HLA-A, and a panel of transfectants expressing the products of A2.1 genes that had been mutated at multiple positions encoding residues in the 2 domain -helix. As a means of detecting distant conformational effects, these altered A2.1 molecules were also examined serologically. Amino acid substitutions at the carboxy-terminal end of the 2 domain -helix led to diminished staining with the monoclonal antibody (mAb) MA2.1. The epitope for this antibody has previously been mapped to the 1 domain -helix (residues 62–65). This suggests that interdomain contacts may cause conformational alteration, and that mutants can have distant, as well as local effects. Of the 24 positions where substitutions were made, only six led to loss of the anti-A2 alloresponse by the three clones and three lines that were tested. In addition, the mutations that altered the MA2.1 epitope, located on the 1 domain -helix, did not inhibit allorecognition. This suggests that a limited number of regions on the A2.1 molecule are responsible for allodeterminant expression. The most influential substitutions were those at positions 152, 154, 162, and 166. It is notable that three of these are predicted to be T-cell receptor (Tcr)-contacting residues, and one (152) to contribute to peptide binding. These results suggest that the specificity of alloreactive T cells is determined by exposed polymorphisms, directly contacted by the Tcr, and by concealed polymorphisms which influence peptide binding.  相似文献   

10.
Glucoamylases from Aspergillus niger and Rhizopus niveus catalyzed condensation of 2-deoxy-D-glucose (dGlc) to yield deoxy-glucooligosaccharides with polymerization degrees of 2–5. The enzymes also gave a small amount of products from 3-deoxy-o-glucose, but no products from 6-deoxy-D-glucose. A. niger α-glucosidase also catalyzed condensation of dGlc, while Torula and Saccharomyces α-glucosidases had low activity. α-l,4-, 1,6-, and 1,3-linked deoxy-glucobioses were isolated and identified as the products of A. niger glucoamylase and A. niger α-glucosidase. In the reaction of the glucoamylase, 1,4- and 1,3-linked saccharides decreased with an increase of 1,6-linked one. A. niger α-glucosidase produced α-1,6-linked disaccharide predominantly during the whole course of the reaction.  相似文献   

11.
12.
The use of prostaglandins E2 and F2α, administered by extra-amniotic instillation, for the induction of abortion was studied in 94 patients in the first and second trimesters of pregnancy. Abortion was successfully induced in 87% of patients within 36 hours and in 94% within 48 hours. The mean abortion time was 22·4 hours. In 60% of patients abortion was complete.Though the differences were not statistically significant, on average multigravid patients aborted more quickly than primigravidae, while the mean abortion time in PGE2-treated patients was less than in those receiving PGF2α.No serious complications occurred. Some side effects were observed. Occasional vomiting was the commonest symptom but the incidence of side effects was lower than with alternative routes of administration. A leucocytosis was often noted but there were no significant instances of infection.The method has proved a safe and effective means of terminating pregnancies in the second trimester.  相似文献   

13.
The degree of hydroxylation of the lysine residue located in both alpha(1)- and alpha(2)-chains of collagen in the N-terminal, non-helical telopeptide region of the molecule has been determined in collagen from various sources after isolation of the peptides (alpha(1)- and alpha(2)-CB1) that contain the lysine residue in question and are obtained by cyanogen bromide cleavage of collagen alpha(1)- and alpha(2)-chains respectively. As with collagen from chick tibia, bone collagens from rat tibia and femur and embryonic chick frontal bone, have a high degree of hydroxylation (approx. 50% or more) of the lysine residue in both alpha(1)- and alpha(2)-CB1 peptides. This is in contrast with the lack of hydroxylation of this residue in both alpha(1)- and alpha(2)-chains of all skin collagens so far examined. The presence of hydroxylysine in alpha(1)- and alpha(2)-CB1 peptides from tendon collagen is also indicated. In rat tail tendon collagen the amount of hydroxylation is only slight but in the much less soluble tendon collagen from embryonic chick leg tendons, approximately one-third of the lysine is hydroxylated.  相似文献   

14.
Abstract

Terazosin and its enantiomers, antagonists of α1-adrenoceptors, were studied in radioligand binding and functional assays to determine relative potencies at subtypes of α1- and α2-adrenoceptors in vitro. The racemic compound and its enantiomers showed high and apparently equal affinity for subtypes of α1-adrenoceptors with K values in the low nanomolar range, and showed potent antagonism of α1-adrenoceptors in isolated tissues, with the enantiomers approximately equipotent to the racemate at each α1-adrenoceptor subtype. At α2b sites, R(+) terazosin bound less potently than either the S(-) enantiomer or racemate. R(+) terazosin was also less potent than the S(-) enantiomer or the racemate at rat atrial α2B receptors. These agents were not significantly different in their potencies at α2a or α2A sites. Since the high affinity for α2B sites of quinazoline-type α-adrenoceptor antagonists has been used to differentiate α2-adrenoceptor subtypes, the low affinity of R(+) terazosin for these sites was unexpected. Because terazosin or its enantiomers are approximately equipotent at α1 -adrenoceptor subtypes, the lower potency of R(+) terazosin at α2B receptors indicates a somewhat greater selectivity for α1- compared to α2B adrenoceptor subtypes. The possible pharmacological significance of this observation is discussed.  相似文献   

15.
Peptide T-11, a carboxyl terminal tryptic fragment of α2-plasmin inhibitor, inhibits the reversible first step of the reaction between plasmin and α2-plasmin inhibitor. To elucidate which amino-acid residues played a important role in the inhibitory activity of peptide T-11, we prepared the various synthetic derivatives of peptide T-11 and determined the peptide concentration that inhibited the apparent rate constant of the reaction between plasmin and α2-plasmin inhibitor by 50% (IC50). Peptide III, which lacked the residues Gly-1 to Pro-7 of peptide I (peptide T-11), had a strong inhibitory activity, like peptide I (IC50: peptide 1, 7 μM; peptide III, 13 μM). The peptides that lacked the Leu-9 and Lys-10 or Lys-26 of peptide III showed much weaker activity, and the loss of amidation of the C-terminal lysine of peptide III also markedly reduced the inhibitory activity, Peptide III competitivef inhibited the binding of [14C]tranexamic acid to kringle 1 + 2 + 3 (K1–3) and kringle 4 (K4) in a binding assay performed by the gel-diffusion method. The respectively dissociation constants (Kd) of peptide III for K1–3 and K4 were 0.85 μM and 35.2 μM. These data suggest that the amino residue of Lys-10 and the carboxylic acid of Lys-26 in peptide T-11 play crucial roles in the ionic binding of α2-plasmin inhibitor to the tranexamic acid-binding site (lysine-binding site) of plasminogen. Peptide T-11: H-G-D-K-L-F-G-P-D-L-K-L-V-P-P-M-E-E-D-Y-P-Q-F-G-S-P-K-OH.  相似文献   

16.
alpha-Heterocyclic alpha-aminoesters were obtained in good yields by reaction of a glycine cation equivalent and different heterocyclic nucleophiles; diastereoselectivity using a carbohydrate (galactopyranose) as N-protecting group was modest.  相似文献   

17.
Organic solvent-resistant Aspergillus niger α-glucosidase (ANGase) can synthesize α-2-deoxyglucosyl derivatives (2DDs) in water-organic solvent media by a trans-addition reaction from d-glucal to various acceptors. Herein, we studied the influence of four different solvents on ANGase stability and activity. ANGase exhibited 47 or 43% residual activity following incubation in 50% (v/v) or in 70% (v/v) acetone for 4 h, respectively. When various carbohydrates were used as acceptor molecules, ANGase catalyzed the addition reaction of four different sugar alcohols, glucose, sucrose, or trehalose to d-glucal. Among the acceptor molecules tested, xylitol was the best acceptor by producing the highest yield (87% addition). The concentration of acetone/acceptor influenced the formation of 2DDs and the yields. We confirmed the molecular weight of five kinds of products by mass spectrometry and enzymatic hydrolysis. Current method is useful for the production of carbohydrates containing 2-deoxyglucose moiety.  相似文献   

18.
The unique steric inhibition of endopeptidases by human alpha(2)M (alpha(2)-macroglobulin) and the inactivation of the latter by methylamine were examined in relation to each other. Progressive binding of trypsin by alpha(2)M was closely correlated with the loss of the methylamine-reactive sites in alpha(2)M: for each trypsin molecule bound, two such sites were inactivated. The results further showed that, even at low proteinase/alpha(2)M ratios, no unaccounted loss of trypsin-binding capacity occurred. As alpha(2)M is bivalent for trypsin binding and no trypsin bound to electrophoretic slow-form alpha(2)M was observed, this indicates that the two sites must react (bind trypsin) in rapid succession. Reaction of [(14)C]methylamine with alpha(2)M was biphasic in time; in the initial rapid phase complex-formation with trypsin caused a largely increased incorporation of methylamine. In the subsequent slow phase trypsin had no such effect. These results prompted further studies on the kinetics of methylamine inactivation of alpha(2)M with time of methylamine treatment. It was found that conformational change of alpha(2)M and decrease in trypsin binding (activity resistant to soya-bean trypsin inhibitor) showed different kinetics. The latter decreased rapidly, following pseudo-first-order kinetics. Conformational change was much slower and followed complex kinetics. On the other hand, binding of (125)I-labelled trypsin to alpha(2)M did follow the same kinetics as the conformational change. This discrepancy between total binding ((125)I radioactivity) and trypsin-inhibitor-resistant binding of trypsin indicated formation of anomalous complexes, in which trypsin could still be inhibited by soya-bean trypsin inhibitor. Further examination confirmed that these complexes were proteolytically active towards haemoglobin and bound (125)I-labelled soya-bean trypsin inhibitor to the active site of trypsin. The inhibition by soya-bean trypsin inhibitor was slowed down as compared with reaction with free trypsin. The results are discussed in relation to the subunit structure of alpha(2)M and to the mechanism of formation of the complex.  相似文献   

19.
Mouse class I antigens are the major targets of cytolytic T lymphocytes in both major histocompatibility complex (MHC)-restricted and allogeneic responses. Considerable evidence has recently accumulated demonstrating that MHC class I molecules encoded by genes whose 1 and 2 coding exons were interchanged are not recognized by T lymphocytes specific for parental class I products. Along with the loss of T -cell reactivity, there is a loss of recognition by some, but not all monoclonal antibodies. In this communication we report that the loss of reactivity by monoclonal antibodies is accompanied by the gain of new epitopes caused by the interaction of 1 and 2 domains. These epitopes are immunodominant. They are the major determinant recognized by polyclonal antisera raised by immunization with L cells transfected with exon-shuffled class I genes. Four new monoclonal antibodies have been produced which recognize at least two separate epitopes caused by the interaction of the 1P and 2d domains.  相似文献   

20.
Photo-oxygenation of α-ionone was studied to clarify the relationship between the maturity of aroma and photo-oxygenative change of α-ionone. α-Ionone was converted to oxygenated derivatives which were identified as 2,3-epoxy-β-ionone, 3,4-epoxy-α-ionone, 4-keto-β-ionone (trans- and cis-form), 5-keto-α-ionone and 3,4-dihydroxy-α-ionone.  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