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.
Histidyl-tRNA synthetase (HisRS) is responsible for the synthesis of histidyl-transfer RNA, which is essential for the incorporation of histidine into proteins. This amino acid has uniquely moderate basic properties and is an important group in many catalytic functions of enzymes. A compilation of currently known primary structures of HisRS shows that the subunits of these homo-dimeric enzymes consist of 420-550 amino acid residues. This represents a relatively short chain length among aminoacyl-tRNA synthetases (aaRS), whose peptide chain sizes range from about 300 to 1100 amino acid residues. The crystal structures of HisRS from two organisms and their complexes with histidine, histidyl-adenylate and histidinol with ATP have been solved. HisRS from Escherichia coli and Thermus thermophilus are very similar dimeric enzymes consisting of three domains: the N-terminal catalytic domain containing the six-stranded antiparallel beta-sheet and the three motifs characteristic of class II aaRS, a HisRS-specific helical domain inserted between motifs 2 and 3 that may contact the acceptor stem of the tRNA, and a C-terminal alpha/beta domain that may be involved in the recognition of the anticodon stem and loop of tRNA(His). The aminoacylation reaction follows the standard two-step mechanism. HisRS also belongs to the group of aaRS that can rapidly synthesize diadenosine tetraphosphate, a compound that is suspected to be involved in several regulatory mechanisms of cell metabolism. Many analogs of histidine have been tested for their properties as substrates or inhibitors of HisRS, leading to the elucidation of structure-activity relationships concerning configuration, importance of the carboxy and amino group, and the nature of the side chain. HisRS has been found to act as a particularly important antigen in autoimmune diseases such as rheumatic arthritis or myositis. Successful attempts have been made to identify epitopes responsible for the complexation with such auto-antibodies.  相似文献   

2.
The accuracy of protein synthesis relies on the ability of aminoacyl-tRNA synthetases (aaRSs) to discriminate among true and near cognate substrates. To date, analysis of aaRSs function, including identification of residues of aaRS participating in amino acid and tRNA discrimination, has largely relied on the steady state kinetic pyrophosphate exchange and aminoacylation assays. Pre-steady state kinetic studies investigating a more limited set of aaRS systems have also been undertaken to assess the energetic contributions of individual enzyme-substrate interactions, particularly in the adenylation half reaction. More recently, a renewed interest in the use of rapid kinetics approaches for aaRSs has led to their application to several new aaRS systems, resulting in the identification of mechanistic differences that distinguish the two structurally distinct aaRS classes. Here, we review the techniques for thermodynamic and kinetic analysis of aaRS function. Following a brief survey of methods for the preparation of materials and for steady state kinetic analysis, this review will describe pre-steady state kinetic methods employing rapid quench and stopped-flow fluorescence for analysis of the activation and aminoacyl transfer reactions. Application of these methods to any aaRS system allows the investigator to derive detailed kinetic mechanisms for the activation and aminoacyl transfer reactions, permitting issues of substrate specificity, stereochemical mechanism, and inhibitor interaction to be addressed in a rigorous and quantitative fashion.  相似文献   

3.
The rate of transfer of amino acid from enzyme-bound aminoacyl adenylate to tRNA has been compared with the rate of esterification of free amino acid. The approach of L?vgren et al. (L?vgren, T. N. E., Heinonen, J., and Loftfield, R. B. (1975) J. Biol. Chem. 250, 3854-3860) was used, with 14C in the aminoacyl adenylate and 3H in the free amino acid and with both the lysine and isoleucine systems of Escherichia coli. In both systems kinetic analyses show more rapid transfer from the preformed enzyme complex when interference by the back reaction with inorganic pyrophosphate was eliminated. Parallel experiments, in which the amount of enzyme complex was measured, confirmed that aminoacyl adenylate is an intermediate in both systems. No evidence was found for an alternative mechanism.  相似文献   

4.
氨基酰tRNA合成酶(aminoacyl-tRNA synthetases,aaRS)家族的经典功能是催化氨基酸与对应tRNA结合,形成氨基酰tRNA,参与蛋白质合成。aaRS在进化过程中不断增加与氨基酰化功能无关的新结构域,其亚细胞器定位也受到营养、压力信号、参与调控血管新生和炎症反应等内外部信号调控,且不同aaRS的突变导致不同人类疾病,提示aaRS具有信号传导功能,但缺少具体的生化机制。最新发现aaRS具有氨基酰转移酶活性。一种氨基酸可以被其对应的aaRS活化成氨基酰AMP,氨基酰AMP可以修饰与该aaRS相互作用蛋白质的赖氨酸,传递该氨基酸的丰度及结构信息,调控细胞信号网络。aaRS新功能的发现和研究,为解释aaRS的生理病理重要性提供新的方向。本文综述aaRS的进化及非经典功能,讨论aaRS氨基酰转移酶活性在细胞信号传导及其与疾病的相关性,也包括药物开发潜力。  相似文献   

5.
Total RNA as well as transfer RNA were quantified from mature ova apart from four different embryonic stages namely mid-cleavage, early gastrula, mid-gastrula and organogenesis of the freshwater teleostHeteropneustes fossilis. Total RNA as well as transfer RNA quantity follow a similar variation pattern, being maximum during mid-gastrulation. When analysed by total amino acid acceptance capacity, transfer RNA shows its maximum activity during mid-gastrulation. This coincides with the higher ratio of tRNA to total RNA at this stage. The relative aminoacylation capacity for Ser, Gly, Asn and Thr are found to be higher (9–34%) compared to that for other amino acids. Total tRNA, resolved into three peaks upon HPLC fractionation, shows a high cumulative peak area during mid-gastrulation and organogenesis. These results indicate a switch over of maternal to embryonic translation machinery during gastrulation.Abbreviations tRNA transfer RNA - aaRS aminoacyl tRNA synthetase - HPLC high pressure liquid chromatography - AUF absorption unit full scale  相似文献   

6.
The shape of the time curve for the aminoacylation of tRNA has been investigated using five different amino acid:tRNA ligases. Four of these enzymes showed a lag in the time curve during the early phase of the first catalytic turnover of the enzyme. In each case, the lag period could be abolished by preincubating the ligase with amino acid, ATP, and Mg2+ under conditions known to give an aminoacyl adenylate-enzyme complex. With all five ligases the steady state rate of transfer from the preformed aminoacyl-adenylate complex to tRNA was approximately the same as that of the overall reaction.  相似文献   

7.
A widespread consensus holds that protein synthesis according to a genetic code was launched entirely by sophisticated RNA molecules that played both coding and functional roles. This belief persists, unsupported by phylogenetic evidence for ancestral ribozymes that catalyzed either amino acid activation or tRNA aminoacylation. By contrast, we have adduced strong experimental evidence that the most highly conserved portions of contemporary aminoacyl-tRNA synthetases (aaRS) accelerate both reactions well in excess of rates achieved by RNA aptomers derived from combinatorial libraries and of rates required for primordial protein synthesis. Such ancestral enzymes, or “Urzymes”, characterized for Class I (TrpRS (Pham et al., 2010, 2007) and LeuRS (Collier et al., 2013); 130 residues) and Class II (HisRS; 120–140 residues; (Li et al., 2011)) synthetases generally have promiscuous amino acid specificities, whereas ATP and cognate tRNA affinities are within an order of magnitude of those for contemporary enzymes. These characteristics match or exceed expectations for the primordial catalysts necessary to launch protein synthesis. Structural hierarchies in Class I and II aaRS also exhibit plateaus of increasing enzymatic activity, suggesting that catalysis by peptides similar to the Aleph motif identified by Trifonov (Sobolevsky et al.) may have been both necessary and sufficient to launch protein synthesis. Sense/antisense alignments of TrpRS and HisRS Urzyme coding sequences reveal unexpectedly high middle-base complementarity that increases in reconstructed ancestral nodes (Chandrasekaran et al.), consistent with the proposal of Rodin and Ohno (Rodin & Ohno, 1995). Thus, these ancestors were likely coded by opposite strands of the same gene, favoring simultaneous expression of aaRS activating both hydrophobic (core) and hydrophilic (surface) amino acids. Our results support the view that aaRS coevolved with cognate tRNAs from a much earlier stage than that envisioned under the RNA World hypothesis, and that their descendants make up appreciable portions of the proteome.  相似文献   

8.
The contemporary genetic code and the process of protein biosynthesis most assuredly evolved from a simpler code and process. We believe that there was obligatory coevolution of the two and that the earlier code and process must have involved a more direct linkage between the amino acids and the information macromolecule. We propose that an early form of translating existed in which amino acids were attached directly to the 'messenger' RNA along the backbone as 2'OH aminoacyl esters. These esters then condensed with each other on the RNA backbone yielding a peptide covalently attached to the RNA, without the use of tRNA's and ribosomes. THis presentation is concerned with experimental data which indicate that such a simple translation system is possible and must have involved the following steps: (1) formation of the aminoacyl adenylate anhydride, (2) transfer of the amino acid from the adenylate to immidazole, (3) transfer of the amino acid from imidazole to 2'OH groups along the backbone of RNAs, (4) condensation of the amino acids to yield peptides. Steps (1)-(3) have been confirmed in chemical systems. Our preliminary evidence indicates step (4) is also possible. The aminoacylation of polyribonucleotides and the subsequent formation of peptides is a dynamic and experimentally accessible system for studying genetic coding specfities and our present studies are now concentrated on step (4), looking for such specifities.  相似文献   

9.
Transfer RNA was analyzed qualitatively as well as quantitatively from ovaries of the fresh water teleostHeteropneustes fossilis for twelve months. The tRNA samples were found to be pure and devoid of any high molecular weight RNA or DNA contaminations. The quantity of tRNA as well as its biological activity, assayed byin vitro aminoacylation using homologous aminoacyl tRNA synthetases, were found to be higher during resting and preparatory (pre-vitellogenic) phases, i.e. from November to March, as compared to vitellogenic and spawning phases of the fish, i.e. from April to October. The highest tRNA pool and its activity was found in the month of February, which coincides with the early preparatory phase. The results indicate that the accumulation of active tRNA starts in the resting phase. Such an accumulation of tRNA may be a part of the enrichment of mature eggs with complete translational machinery before ovulation in order to cope with the high rate of protein synthesis after fertilization.Abbreviations aaRS aminoacyl tRNA synthetase - [14C] APH [14C]-algal protein hydrolysate - ATP adenosine triphosphate - DTT dithiothreitol - EDTA ethylene diamine tetra acetic acid - GSI gonado somatic index - TCA trichloroacetic acid - tRNA transfer RNA  相似文献   

10.
Aminoacyl-tRNA synthetases (ARSs) join amino acids to their cognate tRNAs to initiate protein synthesis. Class II ARS possess a unique catalytic domain fold, possess active site signature sequences, and are dimers or tetramers. The dimeric class I enzymes, notably TyrRS, exhibit half-of-sites reactivity, but its mechanistic basis is unclear. In class II histidyl-tRNA synthetase (HisRS), amino acid activation occurs at different rates in the two active sites when tRNA is absent, but half-of-sites reactivity has not been observed. To investigate the mechanistic basis of the asymmetry, and explore the relationship between adenylate formation and conformational events in HisRS, a fluorescently labeled version of the enzyme was developed by conjugating 7-diethylamino-3-((((2-maleimidyl)ethyl)amino)carbonyl)coumarin (MDCC) to a cysteine introduced at residue 212, located in the insertion domain. The binding of the substrates histidine, ATP, and 5′-O-[N-(l-histidyl)sulfamoyl]adenosine to MDCC-HisRS produced fluorescence quenches on the order of 6–15%, allowing equilibrium dissociation constants to be measured. The rates of adenylate formation measured by rapid quench and domain closure as measured by stopped-flow fluorescence were similar and asymmetric with respect to the two active sites of the dimer, indicating that conformational change may be rate-limiting for product formation. Fluorescence resonance energy transfer experiments employing differential labeling of the two monomers in the dimer suggested that rigid body rotation of the insertion domain accompanies adenylate formation. The results support an alternating site model for catalysis in HisRS that may prove to be common to other class II aminoacyl-tRNA synthetases.The aminoacyl-tRNA synthetases (ARSs)2 comprise two distinct classes of enzymes, all of which catalyze a two-step reaction to generate aminoacyl-tRNA for protein synthesis (1, 2) (Reactions 1 and 2). During the first of two partial reactions in aminoacylation, the cognate amino acid is condensed with ATP to form an aminoacyl-adenylate. This half reaction proceeds by an associative mechanism in which the stereochemistry of the α-phosphate undergoes inversion (3). The adenylate then undergoes a subsequent attack by the cognate tRNA, with the amino acid undergoing transfer to the 3′-terminal adenosine. Aminoacyl transfer requires the activation of 2′ or 3′ of the terminal hydroxyl, and its rate may be accelerated by a number of different mechanisms, including proton transfer to the adenylate, and proton shuttling to the 2′-OH and then to neighboring active site residues (4, 5). Many ARSs can activate their cognate amino acids in the absence of tRNA, allowing the two partial reactions to be studied individually. Notably, there are significant gaps in our understanding of how the adenylation and aminoacyl transfer half reactions are integrated into the overall reaction schemes of ARSs.Class I and class II enzymes can be broadly distinguished by their oligomeric structure. The former are generally monomeric, whereas the latter are typically dimeric or tetrameric (6). Notable exceptions to this pattern are the class Ic tyrosyl- and tryptophanyl-tRNA synthetases, both of which form obligatory dimers (7, 8). Both have been described as possessing half-of-sites reactivity (9, 10), but the picture is more complex. Consistent with half-of-sites reactivity, TyrRS binds one mole of tyrosine per dimer and retains a single mole of adenylate per mole of dimers when the E·Ade complex is purified away from unreacted substrates by size-exclusion chromatography (11). However, the steady-state kinetics of TyrRS show no evidence of cooperativity, the second binding site becomes accessible to substrates when the first site is occupied by adenylate, and TyrRS clearly binds 2 mol of tRNA in the crystal (7, 12).On the basis of these and other observations involving the rate of hydrolysis of the on-enzyme adenylate, Fersht (13) proposed that the second site of TyrRS possesses weak catalytic activity and that TyrRS is asymmetric in solution. The impact of this potential asymmetry in the activation reaction on the complete TyrRS catalytic cycle remains to be explored. TrpRS also exhibits half-of-sites reactivity, and a detailed analysis of the aminoacyl transfer reaction by pre-steady state kinetics proposed both random and ordered versions of alternating site catalysis as models of the enzyme (14). In the class II ARSs, the tetrameric SepRS represents the single example where half-of-sites reactivity has been demonstrated experimentally (15).Despite this apparent class distinction, recent work on HisRS, a class IIa ARS that is well characterized with respect to structure (1619), tRNA recognition, and reaction kinetics (4, 20), highlighted several functional attributes that are reminiscent of class I TyrRS. Like TyrRS, HisRS retains only 1 mol of adenylate per dimer when subjected to size-exclusion chromatography (4). A detailed pre-steady-state analysis of mutants of tRNAHis or HisRS compromised with respect to tRNA identity suggested that, in the complete aminoacylation reaction, formation of aminoacyl adenylate in the second active site is contingent upon a productive aminoacyl transfer reaction in the first (20). These and other data led to the proposal of an alternating site model for HisRS (20) that is analogous to the “flip flop” catalysis suggested for class II PheRS (21, 22) and class Ic TrpRS (14). This raises the possibility that the catalytic cycles of dimeric class II enzymes and dimeric class Ic enzymes share some common feature.Alternating catalysis requires a mechanism for coupling events between active sites, presumably through conformational changes propagated between these active sites. To investigate these events, a version of HisRS was developed that featured the site-specific incorporation of extrinsic environmentally sensitive fluorescent probes, allowing the adenylation reaction to be followed by stopped-flow fluorometry. Comparison of the kinetics of substrate-induced fluorescence changes to the kinetics of product formation determined by rapid quench suggests that adenylation rates are asymmetric with respect to the two active sites of the dimer, and that conformational changes linked to the insertion domain may be rate-limiting for product formation. The implications of these results for a previous model (20) of alternating site catalysis in HisRS are discussed.  相似文献   

11.
Glutaminyl-tRNA synthetase generates Gln-tRNA(Gln) 10(7)-fold more efficiently than Glu-tRNA(Gln) and requires tRNA to synthesize the activated aminoacyl adenylate in the first step of the reaction. To examine the role of tRNA in amino acid activation more closely, several assays employing a tRNA analog in which the 2'-OH group at the 3'-terminal A76 nucleotide is replaced with hydrogen (tRNA(2'HGln)) were developed. These experiments revealed a 10(4)-fold reduction in kcat/Km in the presence of the analog, suggesting a direct catalytic role for tRNA in the activation reaction. The catalytic importance of the A76 2'-OH group in aminoacylation mirrors a similar role for this moiety that has recently been demonstrated during peptidyl transfer on the ribosome. Unexpectedly, tracking of Gln-AMP formation utilizing an alpha-32P-labeled ATP substrate in the presence of tRNA(2'HGln) showed that AMP accumulates 5-fold more rapidly than Gln-AMP. A cold-trapping experiment revealed that the nonenzymatic rate of Gln-AMP hydrolysis is too slow to account for the rapid AMP formation; hence, the hydrolysis of Gln-AMP to form glutamine and AMP must be directly catalyzed by the GlnRS x tRNA(2'HGln) complex. This hydrolysis of glutaminyl adenylate represents a novel reaction that is directly analogous to the pre-transfer editing hydrolysis of noncognate aminoacyl adenylates by editing synthetases such as isoleucyl-tRNA synthetase. Because glutaminyl-tRNA synthetase does not possess a spatially separate editing domain, these data demonstrate that a pre-transfer editing-like reaction can occur within the synthetic site of a class I tRNA synthetase.  相似文献   

12.
Abstract

The genetic code is based on the aminoacylation of tRNA with amino acids catalyzed by the aminoacyl-tRNA synthetases. The synthetases are constructed from discrete domains and all synthetases possess a core catalytic domain that catalyzes amino acid activation, binds the acceptor stem of tRNA, and transfers the amino acid to tRNA. Fused to the core domain are additional domains that mediate RNA interactions distal to the acceptor stem. Several synthetases catalyze the aminoacylation of RNA oligonucleotide substrates that recreate only the tRNA acceptor stems. In one case, a relatively small catalytic domain catalyzes the aminoacylation of these substrates independent of the rest of the protein. Thus, the active site domain may represent a primordial synthetase in which polypeptide insertions that mediate RNA acceptor stem interactions are tightly integrated with determinants for aminoacyl adenylate synthesis. The relationship between nucleotide sequences in small RNA oligonucleotides and the specific amino acids that are attached to these oligonucleotides could constitute a second genetic code.  相似文献   

13.
RNA minihelices and the decoding of genetic information   总被引:1,自引:0,他引:1  
P Schimmel 《FASEB journal》1991,5(8):2180-2187
The rules of the genetic code are determined by the specific aminoacylation of transfer RNAs by aminoacyl transfer RNA synthetase. A straightforward analysis shows that a system of synthetase-tRNA interactions that relies on anticodons for specificity could, in principle, enable most synthetases to distinguish their cognate tRNA isoacceptors from all others. Although the anticodons of some tRNAs are recognition sites for the cognate aminoacyl tRNA synthetases, for other synthetases the anticodon is dispensable for specific aminoacylation. In particular, alanine and histidine tRNA synthetases aminoacylate small RNA minihelices that reconstruct the part of their cognate tRNAs that is proximate to the amino acid attachment site. Helices with as few as six base pairs can be efficiently aminoacylated. The specificity of aminoacylation is determined by a few nucleotides and can be converted from one amino acid to another by the change of only a few nucleotides. These findings suggest that, for a subgroup of the synthetases, there is a distinct code in the acceptor helix of transfer RNAs that determines aminoacylation specificity.  相似文献   

14.
Nordin BE  Schimmel P 《Biochemistry》2003,42(44):12989-12997
The genetic code depends on amino acid fine structure discrimination by aminoacyl-tRNA synthetases. For isoleucyl- (IleRS) and valyl-tRNA synthetases (ValRS), reactions that hydrolyze misactivated noncognate amino acids help to achieve high accuracy in aminoacylation. Two editing pathways contribute to aminoacylation fidelity: pretransfer and post-transfer. In pretransfer editing, the misactivated amino acid is hydrolyzed as an aminoacyl adenylate, while in post-transfer editing a misacylated tRNA is deacylated. Both reactions are dependent on a tRNA cofactor and require translocation to a site located approximately 30 A from the site of amino acid activation. Using a series of 3'-end modified tRNAs that are deficient in either aminoacylation, deacylation, or both, total editing (the sum of pre- and post-transfer editing) was shown to require both aminoacylation and deacylation activities. These and additional results with IleRS are consistent with a post-transfer deacylation event initiating formation of an editing-active enzyme/tRNA complex. In this state, the primed complex processively edits misactivated valyl-adenylate via the pretransfer route. Thus, misacylated tRNA is an obligatory intermediate for editing by either pathway.  相似文献   

15.
The contemporary genetic code and the process of protein biosynthesis most assuredly evolved from a simpler code and process. We believe that there was obligatory coevolution of the two and that the earlier code and process must have involved a more direct linkage between the amino acids and the informational macromolecule. We propose that an early form of translating existed in which amino acids were attached directly to the ‘messenger’ RNA along the backbone as 2'OH aminoacyl esters. These esters then condensed with each other on the RNA backbone yielding a peptide covalently attached to the RNA, without the use of tRNAs and ribosomes. This presentation is concerned with experimental data which indicate that such a simple translation system is possible and must have involved the following steps: (1) formation of the aminoacyl adenylate anhydride, (2) transfer of the amino acid from the adenylate to imidazole, (3) transfer of the amino acid from imidazole to 2'OH groups along the backbone of RNAs (4) condensation of the amino acids to yield peptides. Steps (1)–(3) have been confirmed in chemical systems. Our preliminary evidence indicates step (4) is also possible. The aminoacylation of polyribonucleotides and the subsequent formation of peptides is a dynamic and experimentally accessible system for studying genetic coding specifities and our present studies are now concentrated on step (4), looking for such specifities.  相似文献   

16.
A J Lloyd  H U Thomann  M Ibba    D Sll 《Nucleic acids research》1995,23(15):2886-2892
We describe a convenient, simple and novel continuous spectrophotometric method for the determination of aminoacyl-tRNA synthetase activity. The assay relies upon the measurement of inorganic pyrophosphate generated in the first step of the aminoacylation of a tRNA. Pyrophosphate release is coupled to inorganic pyrophosphatase, to generate phosphate, which in turn is used as the substrate of purine nucleoside phosphorylase to catalyze the N-glycosidic cleavage of 2-amino 6-mercapto 7-methylpurine ribonucleoside. Of the reaction products, ribose 1-phosphate and 2-amino 6-mercapto 7-methylpurine, the latter has a high absorbance at 360 nm relative to the nucleoside and hence provides a spectrophotometric signal that can be continuously followed. The non-destructive nature of the spectrophotometric assay allowed the re-use of the tRNAs in question in successive experiments. The usefulness of this method was demonstrated for glutaminyl-tRNA synthetase (GlnRS) and tryptophanyl-tRNA synthetase. Initial velocities measured using this assay correlate closely with those assayed by quantitation of [3H]Gln-tRNA or [14C]Trp-tRNA formation respectively. In both cases amino acid transfer from the aminoacyl adenylate to the tRNA represents the rate determining step. In addition, aminoacyl adenylate formation by aspartyl-tRNA synthetase was followed and provided a more sensitive means of active site titration than existing techniques. Finally, this novel method was used to provide direct evidence for the cooperativity of tRNA and ATP binding to GlnRS.  相似文献   

17.
Farrow MA  Schimmel P 《Biochemistry》2001,40(14):4478-4483
Aminoacyl-tRNA synthetases establish the rules of the genetic code by aminoacylation reactions. Occasional activation of the wrong amino acid can lead to errors of protein synthesis. For isoleucyl-tRNA synthetase, these errors are reduced by tRNA-dependent hydrolytic editing reactions that occur at a site 25 A from the active site. These reactions require that the misactivated amino acid be translocated from the active site to the center for editing. One mechanism describes translocation as requiring the mischarging of tRNA followed by a conformational change in the tRNA that moves the amino acid from one site to the other. Here a specific DNA aptamer is investigated. The aptamer can stimulate amino acid-specific editing but cannot be aminoacylated. Although the aptamer could in principle stimulate hydrolysis of a misactivated amino acid by an idiosyncratic mechanism, the aptamer is shown here to induce translocation and hydrolysis of misactivated aminoacyl adenylate at the same site as that seen with the tRNA cofactor. Thus, translocation to the site for editing does not require joining of the amino acid to the nucleic acid. Further experiments demonstrated that aptamer-induced editing is sensitive to aptamer sequence and that the aptamer is directed to a site other than the active site or tRNA binding site of the enzyme.  相似文献   

18.
In the present work we report, for the first time, a novel difference in the molecular mechanism of the activation step of aminoacylation reaction between the class I and class II aminoacyl tRNA synthetases (aaRSs). The observed difference is in the mode of nucleophilic attack by the oxygen atom of the carboxylic group of the substrate amino acid (AA) to the αP atom of adenosine triphosphate (ATP). The syn oxygen atom of the carboxylic group attacks the α-phosphorous atom (αP) of ATP in all class I aaRSs (except TrpRS) investigated, while the anti oxygen atom attacks in the case of class II aaRSs. The class I aaRSs investigated are GluRS, GlnRS, TyrRS, TrpRS, LeuRS, ValRS, IleRS, CysRS, and MetRS and class II aaRSs investigated are HisRS, LysRS, ProRS, AspRS, AsnRS, AlaRS, GlyRS, PheRS, and ThrRS. The variation of the electron density at bond critical points as a function of the conformation of the attacking oxygen atom measured by the dihedral angle ψ (C(α)-C') conclusively proves this. The result shows that the strength of the interaction of syn oxygen and αP is stronger than the interaction with the anti oxygen for class I aaRSs. This indicates that the syn oxygen is the most probable candidate for the nucleophilic attack in class I aaRSs. The result is further supported by the computation of the variation of the nonbonded interaction energies between αP atom and anti oxygen as well as syn oxygen in class I and II aaRSs, respectively. The difference in mechanism is explained based on the analysis of the electrostatic potential of the AA and ATP which shows that the relative arrangement of the ATP with respect to the AA is opposite in class I and class II aaRSs, which is correlated with the organization of the active site in respective aaRSs. A comparative study of the reaction mechanisms of the activation step in a class I aaRS (Glutaminyl tRNA synthetase) and in a class II aaRS (Histidyl tRNA synthetase) is carried out by the transition state analysis. The atoms in molecule analysis of the interaction between active site residues or ions and substrates are carried out in the reactant state and the transition state. The result shows that the observed novel difference in the mechanism is correlated with the organizations of the active sites of the respective aaRSs. The result has implication in understanding the experimentally observed different modes of tRNA binding in the two classes of aaRSs.  相似文献   

19.
Several analogues of valine, leucine, and isoleucine carrying hydroxyl groups in the gamma- or delta-position have been tested in the aminoacylation of tRNA by valyl-tRNA synthetases from Saccharomyces cerevisiae and Escherichia coli. Results of the ATP/PPi exchange and of the aminoacylation reactions indicate that the amino acid analogues not only can form the aminoacyl adenylate intermediate but are also transferred to tRNA. However, the fact that the reaction consumes an excess of ATP indicates that the misactivated amino acid analogue is hydrolytically removed. Thus, valyl-tRNA synthetase from S. cerevisiae shows a high fidelity in forming valyl-tRNA. Although the much bulkier amino acid analogues allo- and iso-gamma-hydroxyvaline and allo- and iso-gamma-hydroxyisoleucine are initially charged to tRNA, the misaminoacylated tRNA(Val) is enzymatically deacylated. This cleavage reaction is mediated by the hydroxyl groups of the amino acid analogues which are converted into the corresponding lactones.  相似文献   

20.
The fidelity of tRNA aminoacylation is dependent in part on amino acid editing mechanisms. A hydrolytic activity that clears mischarged tRNAs typically resides in an active site on the tRNA synthetase that is distinct from its synthetic aminoacylation active site. A second pre-transfer editing pathway that hydrolyzes the tRNA synthetase aminoacyl adenylate intermediate can also be activated. Pre- and post-transfer editing activities can co-exist within a single tRNA synthetase resulting in a redundancy of fidelity mechanisms. However, in most cases one pathway appears to dominate, but when compromised, the secondary pathway can be activated to suppress tRNA synthetase infidelities.  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