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.
We observed that in vivo and in vitro a small fraction of the glycolytic enzyme enolase became covalently modified by its substrate 2-phosphoglycerate (2-PG). In modified Escherichia coli enolase, 2-PG was bound to Lys341, which is located in the active site. An identical reversible modification was observed with other bacterial enolases, but also with enolase from Saccharomyces cerevisiae and rabbit muscle. An equivalent of Lys341, which plays an important role in catalysis, is present in enolase of all organisms. Covalent binding of 2-PG to this amino acid rendered the enzyme inactive. Replacement of Lys341 of E.coli enolase with other amino acids prevented the automodification and in most cases strongly reduced the activity. As reported for other bacteria, a significant fraction of E.coli enolase was found to be exported into the medium. Interestingly, all Lys341 substitutions prevented not only the automodification, but also the export of enolase. The K341E mutant enolase was almost as active as the wild-type enzyme and therefore allowed us to establish that the loss of enolase export correlates with the loss of modification and not the loss of glycolytic activity.  相似文献   

2.
The nusG gene of Thermus thermophilus HB8 was cloned and sequenced. It is located 388 bp downstream from tufB, which is followed by the genes for ribosomal proteins L11 and L1. No equivalent to secE preceding nusG, as in Escherichia coli, could be detected. The nusG gene product was overproduced in E. coli. A rabbit antiserum raised against the purified recombinant NusG reacted exclusively with one protein band of T. thermophilus crude extracts in Western blot (immunoblot) analyses, and no cross-reaction of the antiserum with E. coli NusG was observed. Recombinant NusG and the reacting T. thermophilus wild-type protein had identical sizes on sodium dodecyl sulfate-polyacrylamide gels. T. thermophilus and E. coli NusG have 45% identical and 22.5% similar amino acids, and similarities between the two proteins are most pronounced in carboxy-terminal regions. The T. thermophilus nusG gene could not rescue a nusG-deficient E. coli mutant strain.  相似文献   

3.
Acetyl-CoA synthetase (ACS) catalyses the activation of acetate to acetyl-CoA in the presence of ATP and CoA. The gene encoding Bradyrhyzobium japonicum ACS has been cloned, sequenced, and expressed in Escherichia coli. The enzyme comprises 648 amino acid residues with a calculated molecular mass of 71,996 Da. The recombinant enzyme was also purified from the transformed E. coli. The enzyme was essentially indistinguishable from the ACS of B. japonicum bacteroids as to the criteria of polyacrylamide gel electrophoresis and biochemical properties. Based on the results of database analysis, Gly-263, Gly-266, Lys-269, and Glu-414 were selected for site-directed mutagenesis in order to identify amino acid residues essential for substrate binding and/or catalysis. Four different mutant enzymes (G263I, G266I, K269G, and E414Q) were prepared and then subjected to steady-state kinetic studies. The kinetic data obtained for the mutants suggest that Gly-266 and Lys-269 participate in the formation of acetyl-AMP, whereas Glu-414 may play a role in acetate binding.  相似文献   

4.
Tryptophan synthase, which catalyzes the final step of tryptophan biosynthesis, is a multifunctional protein that requires pyridoxal phosphate for two of its three distinct enzyme activities. Tryptophan synthase from Neurospora crassa, a homodimer of two 75-kDa subunits, was shown to bind 1 mol of pyridoxal phosphate/mol of subunit with a calculated dissociation constant for pyridoxal phosphate of 1.1 microM. The spectral properties of the holoenzyme, apoenzyme, and reconstituted holoenzyme were characterized and compared to those previously established for the heterotetrameric (alpha 2 beta 2) enzyme from Escherichia coli. The Schiff base formed between pyridoxal phosphate and the enzyme was readily reduced by sodium borohydride, but not sodium cyanoborohydride. The active site residue that binds pyridoxal phosphate, labeled by reduction of the Schiff base with tritium-labeled sodium borohydride, was determined to be lysine by high performance liquid chromatography analysis of the protein hydrolysate. A 5400-dalton peptide containing the reduced pyridoxal phosphate moiety was generated by cyanogen bromide treatment, purified and sequenced. The sequence is 85% homologous with the corresponding sequence obtained for yeast tryptophan synthase (Zalkin, H., and Yanofsky, C. (1982) J. Biol. Chem. 257, 1491-1500); the lysine derivatized by pyridoxal phosphate is located at the same relative position as that in the yeast and E. coli enzymes.  相似文献   

5.
Asp-362, a potential key catalytic residue of Escherichia coli citrate synthase (citrate oxaloacetate-lyase [pro-3S)-CH2COO- ----acetyl-CoA), EC 4.1.3.7) has been converted to Gly-362 by oligonucleotide-directed mutagenesis. The mutant gene was completely sequenced, using a series of synthetic oligodeoxynucleotides spanning the structural gene to confirm that no additional mutations had occurred during genetic manipulation. The mutant gene was expressed in M13 bacteriophage and produced a protein which migrated in an identical manner to wild-type E. coli citrate synthase on SDS-polyacrylamide gels and which cross-reacted with E. coli citrate synthase antiserum. The mutant gene was subsequently recloned into pBR322 for large scale purification of the protein, and the resulting plasmid, pCS31, used to transform the citrate synthase deletion strain, W620. The mutant enzyme purified in an analogous manner to wild-type E. coli citrate synthase and expressed less than 2% of wild-type enzyme activity. The activity of the partial reactions catalysed by citrate synthase was similarly affected suggesting that this residual activity may be due to contaminating wild-type enzyme activity. The mutant citrate synthase retains a high-affinity NADH-binding site consistent with the protein preserving its overall structural integrity. Oxaloacetate binding to the protein is unaffected by the Asp-362 to Gly-362 mutation. Binding of the acetyl-CoA analogue, carboxymethyl-CoA, could not be detected in the mutant protein indicating that the lack of catalytic competence is due primarily to the inability of the protein to bind the second substrate, acetyl-CoA.  相似文献   

6.
H Y Qi  K Sankaran  K Gan    H C Wu 《Journal of bacteriology》1995,177(23):6820-6824
The structure-function relationship of bacterial prolipoprotein diacylgyceryl transferase (LGT) Has been investigated by a comparison of the primary structures of this enzyme in phylogenetically distant bacterial species, analysis of the sequences of mutant enzymes, and specific chemical modification of the Escherichia coli enzyme. A clone containing the gene for LGT, lgt, of the gram-positive species Staphylococcus aureus was isolated by complementation of the temperature-sensitive lgt mutant of E. coli (strain SK634) defective in LGT activity. In vivo and in vitro assays for prolipoprotein diacylglyceryl modification activity indicated that the complementing clone restored the prolipoprotein modification activity in the mutant strain. Sequence determination of the insert DNA revealed an open reading frame of 837 bp encoding a protein of 279 amino acids with a calculated molecular mass of 31.6 kDa. S. aureus LGT showed 24% identity and 47% similarity with E. coli, Salmonella typhimurium, and Haemophilus influenzae LGT.S. aureus LGT, while 12 amino acids shorter than the E. coli enzyme, had a hydropathic profile and a predicted pI (10.4) similar to those of the E. coli enzyme. Multiple sequence alignment among E. coli, S. typhimurium, H. influenzae, and S. aureus LGT proteins revealed regions of highly conserved amino acid sequences throughout the molecule. Three independent lgt mutant alleles from E. coli SK634, SK635, and SK636 and one lgt allele from S. typhimurium SE5221, all defective in LGT activity at the nonpermissive temperature, were cloned by PCR and sequenced. The mutant alleles were found to contain a single base alteration resulting in the substitution of a conserved amino acid. The longest set of identical amino acids without any gap was H-103-GGLIG-108 in LGT from these four microorganisms. In E. coli lgt mutant SK634, Gly-104 in this region was mutated to Ser, and the mutant organism was temperature sensitive in growth and exhibited low LGT activity in vitro. Diethylpyrocarbonate inactivated the E. coli LGT with a second-order rate constant of 18.6 M-1S-1, and the inactivation of LGT activity was reversed by hydroxylamine at pH 7. The inactivation kinetics were consistent with the modification of a single residue, His or Tyr, essential for LGT activity.  相似文献   

7.
L-Aspartate-alpha-decarboxylase, an enzyme that catalyzes the production of beta-alanine, has been purified to apparent homogeneity from Escherichia coli. The properties of the enzyme are: (a) pH optimum of 6.8 to 7.5, (b) temperature optimum of 55 degrees C, (c) Km for L-aspartate of 0.16 mM, and (d) molecular weight of 58,000. The activity of the enzyme is inhibited by reagents (hydroxylamine, phenylhydrazine, and sodium borohydride) that react with carbonyl groups, but no pyridoxal phosphate is present. The compound containing the carbonyl group has been identified as covalently bound pyruvate. Approximately 1 mol of pyruvate was found/mol of enzyme. That the enzyme has a biosynthetic function rather than a catabolic role is indicated by the observations that a mutant (designated as E. coli 99-2) which requires either beta-alanine or pantothenic acid for growth contains only trace amounts of enzyme activity, whereas it is present in substantial amounts in the parent strain (E. coli W) and in a spontaneous revertant of the mutant.  相似文献   

8.
Neutral trehalase was purified from stationary yeast ABYS1 mutant cells deficient in the vacuolar proteinases A and B and the carboxypeptidases Y and S. The purified electrophoretically homogeneous preparation of phosphorylated neutral trehalase exhibited a molecular mass of 160,000 Da on nondenaturing gel electrophoresis and of 80,000 Da on sodium dodecyl sulfate-gel electrophoresis. Maximal activity (114 mumol of trehalose min-1 x mg-1 at 37 degrees C) was observed at pH 6.8-7.0. The apparent Km for trehalose was 34.5 mM. Among seven oligosaccharides studied, the enzyme formed glucose only from trehalose. Neutral trehalase is located in the cytosol. A polyclonal rabbit antiserum raised against neutral trehalase precipitates the enzyme in the presence of protein A. The antiserum does not react with acid trehalase. Dephosphorylation by alkaline phosphatase from Escherichia coli of the active phosphorylated enzyme is accompanied by greater than or equal to 90% inactivation. Rephosphorylation by incubation with the catalytic subunit of beef heart protein kinase is accompanied by reactivation and incorporation of 0.85 mol of phosphate/mol subunit (80,000 Da). The phosphorylated amino acid residue was identified as phosphoserine.  相似文献   

9.
The type I regulatory subunit (R-I) of rat brain cAMP-dependent protein kinase was expressed in E. coli and site-directed mutagenesis was used to substitute amino acids in the putative cAMP-binding sites. The wild-type recombinant R-I bound 2 mol of cAMP/mol subunit, while two mutant R-Is with a single amino acid substitution in one of the two intrachain cAMP-binding sites (clone N153:a glutamate for Gly-200, and clone C254:an aspartate for Gly-324) bound 1 mol of cAMP/mol subunit. When these two substitutions were made in one mutant, cAMP did not bind to this mutant, indicating that binding of cAMP to N153 or C254 was to their nonmutated sites. Competition experiments with site-selective analogs and dissociation of bound cAMP from mutant R-Is provided evidence for strong intrachain interactions between the two classes of cAMP-binding sites in R-I.  相似文献   

10.
The gamma gamma isozyme of rabbit enolase was labeled with fluorescein and the effects of NaClO4 on both enzymatic activity and fluorescence polarization were studied. NaClO4, but not NaCl, dissociates and partially inactivates the enzyme. If dissociation is prevented, either by the addition of substrate or by covalently crosslinking the enzyme, inactivation is also prevented. Analysis of the time and concentration dependence of inactivation and dissociation shows that the decrease in activity is a two-step process: D in equilibrium 2M in equilibrium 2M*. Both monomeric forms of the enzyme are catalytically active.  相似文献   

11.
By site-directed mutagenesis, Thr-75 was converted to Cys-75 in the glutathione reductase (EC 1.6.4.2) of Escherichia coli. This led to the spontaneous formation of an intersubunit disulphide bridge across the 2-fold axis of the dimeric enzyme. The disulphide bridge had no deleterious effect on the catalytic activity, but nor did it increase the thermal stability of the enzyme, possibly because of local conformational flexibility on the dimer interface. The T75C mutant, like the wild-type enzyme, was inactivated by NADPH, proving that this inactivation cannot be due to simple dissociation of the dimer.  相似文献   

12.
The ileS gene encoding the isoleucyl-tRNA synthetase of the thermophilic archaebacterium Methanobacterium thermoautotrophicum Marburg was isolated and sequenced. ileS was closely flanked by an unknown open reading frame and by purL and thus is arranged differently from the organizations observed in several eubacteria or in Saccharomyces cerevisiae. The deduced amino acid sequence of isoleucyl-tRNA synthetase was compared with primary sequences of isoleucyl-, valyl-, leucyl-, and methionyl-tRNA synthetases from eubacteria and yeast. The archaebacterial enzyme fitted well into this group of enzymes. It contained the two short consensus sequences observed in class I aminoacyl-tRNA synthetases as well as regions of homology with enzymes of the isoleucine family. Comparison between the isoleucyl-tRNA synthetases of M. thermoautotrophicum yielded 36% amino acid identity with the yeast enzyme and 32% identity with the corresponding enzyme from Escherichia coli. The ileS gene of the pseudomonic acid-resistant M. thermoautotrophicum mutant MBT10 was also sequenced. The mutant enzyme had undergone a glycine to aspartic acid transition at position 590, in a conserved region comprising the KMSKS consensus sequence. The inhibition constants of pseudomonic acid, KiIle and KiATP, for the mutant enzyme were 10-fold higher than those determined for the wild-type enzyme. Both the mutant and the wild-type ileS gene were expressed in E. coli, and their products displayed the expected difference in sensitivity toward pseudomonic acid.  相似文献   

13.
The correlation between the in vivo functioning and the in vitro behavior of the thermolabile alanyl-transfer ribonucleic acid (tRNA) synthetase (ARS) of Escherichia coli strain BM113 is presented. As a measure for the ARS activity inside the cell, the amount of acylated tRNA(ala) in vivo was determined. The rapid drop of the per cent tRNA(ala) charged which was observed upon shifting a culture of BM113 to the nonpermissive temperature indicates that in vivo acylation of tRNA(ala) might be the growth-limiting step at high temperature. Since neither growth nor the in vivo charging level of tRNA(ala) was affected by the addition of high l-alanine concentrations to the medium, one may infer that impaired functioning of the mutant enzyme at 40 C seems not to be due to reduced affinity of the enzyme for the amino acid. Separation of bulk tRNA of E. coli and of yeast on benzoylated diethylaminoethyl cellulose and charging of the fractions of the column by wild-type and mutant ARS reveal that only those tRNA species aminoacylated by the wild-type enzyme are also charged by the mutant ARS. Determination of the K(m) values of wild-type and mutant ARS for the three isoaccepting tRNA(ala) species of E. coli shows a ca. 10-fold increase of the apparent K(m) values of the mutant enzyme for all three species. Thus, the mutation proportionally reduces the apparent affinity for tRNA(ala) without causing any detectable recognition errors. Investigation of heat inactivation kinetics of wild-type and mutant ARS without and in the presence of substrates provides further evidence that only the transfer site of the ARS is altered by the mutation. Moreover, whereas both enzymes possess the same pH optimum of the relative maximal velocity, their pH dependence of the K(m) values for tRNA is different. The K(m) of the wild-type enzyme decreases at pH values below 7.0 and that of the mutant enzyme shows the inverse tendency; this again indicates an alteration of the tRNA binding site.  相似文献   

14.
The temperature- and solvent-induced denaturation of both the SCP2 wild-type and the mutated protein c71s were studied by CD measurements at 222 nm. The temperature-induced transition curves were deconvoluted according to a two-state mechanism resulting in a transition temperature of 70.5 degrees C and 59.9 degrees C for the wild-type and the c71s, respectively, with corresponding values of the van't Hoff enthalpies of 183 and 164 kJ/mol. Stability parameters characterizing the guanidine hydrochloride denaturation curves were also calculated on the basis of a two-state transition. The transitions of the wild-type occurs at 0.82 M GdnHCl and that of the c71s mutant at 0.55 M GdnHCl. These differences in the half denaturation concentration of GdnHCl reflect already the significant stability differences between the two proteins. A quantitative measure are the Gibbs energies DeltaG(0)(D)(buffer) at 25 degrees C of 15.5 kJ/mol for the wild-type and 8.0 kJ/mol for the mutant. We characterized also the alkyl chain binding properties of the two proteins by measuring the interaction parameters for the complex formation with 1-O-Decanyl-beta-D-glucoside using isothermal titration microcalorimetry. The dissociation constants, K(d), for wild-type SCP2 are 335 microM at 25 degrees C and 1.3 mM at 35 degrees C. The corresponding binding enthalpies, DeltaH(b), are -21. 5 kJ/mol at 25 degrees C and 72.2 kJ/mol at 35 degrees C. The parameters for the c71s mutant at 25 degrees C are K(d)=413 microM and DeltaH(b)=16.6 kJ/mol. These results suggest that both SCP2 wild-type and the c71s mutant bind the hydrophobic compound with moderate affinity.  相似文献   

15.
A recombinant strain of Escherichia coli has been constructed that produces approx. 200 times the amount of hydroxymethylbilane synthase found in wild-type E. coli [Hart, Abell & Battersby (1986) Biochem. J. 240, 273-276]. Enzyme purified from this strain is shown to be permanently inactivated by pyridoxal 5'-phosphate/NaB1H3(3)H1. The inactivation is not complete despite the fact that approx. 1 mol of lysine residues is modified per mol of enzyme. Evidence is gained showing that (a) modification of one of two conserved lysine residues (Lys-55 or Lys-59) results in inactivation of hydroxymethylbilane synthase and (b) these lysine residues are present in or close to the active site.  相似文献   

16.
D A Feldman  P Datta 《Biochemistry》1975,14(8):1760-1767
Incubation of Escherichia coli cells with glucose, pyruvate, and certain other metabolites led to rapid inactivation of inducible biodegradative threonine dehydratase. Analysis with several mutant strains showed that pyruvate, and not a metabolite derived from pyruvate, was capable of inactivating enzyme, and that glucose acted indirectly after being converted to pyruvate. Some other alpha-keto acids such as oxaloacetate and alpha-ketobutyrate (but not alpha-ketoglutarate) were also effective. Inactivation of threonine dehydratase by pyruvate was also observed with purified enzyme preparations. The rates of enzyme inactivation increased with increased concentrations of pyruvate and decreased with increased levels of AMP. Increasing protein concentrations lowered the rates of enzyme inactivation. Dithiothreitol had a large effect on the maximum extent of inactivation of the enzyme by pyruvate; high concentrations of AMP and DTT almost completely counteracted the effect of pyruvate. Gel filtration data showed that pyruvate influenced the oligomeric state of the enzyme by altering the association-dissociation equilibrium in favor of dissociation; the Stokes' radius of the pyruvate-inactivated enzyme was 32 A as compared to 42 A for the untreated enzyme. Reassociation of the dissociated form of the enzyme was achieved by removal of excess free pyruvate by dialysis against buffer supplemented with AMP and DTT. Incubation of threonine dehydratase with [14-C]pyruvate revealed apparent covalent attachment of pyruvate to the enzyme. Strong protein denaturants such as guanidine, urea, and sodium dodecyl sulfate failed to release bound radioactive pyruvate; the molar ratio of firmly bound pyruvate was approximately 1 mol/150,000 g of protein. Pretreatment of the enzyme with p-chloromercuribenzoate and 5,5'-dithiobis(2-nitrobenzoate) (Nbs2) did not reduce the binding of [14-C]pyruvate suggesting no active site SH was involved in the pyruvate-enzyme linkage. Titration of active and pyruvate-inactivated enzyme with Nbs2 indicated that the loss in enzyme activity was not due to oxidation of essential sulfhydryl groups on the enzyme. Based on these data we propose that the mechanism of enzyme inactivation by pyruvate involves covalent attachment of pyruvate to the active oligomeric form of the enzyme followed by dissociation of the oligomer to yield inactive enzyme.  相似文献   

17.
The members of the mechanistically diverse, (beta/alpha)(8)-barrel fold-containing enolase superfamily evolved from a common progenitor but catalyze different reactions using a conserved partial reaction. The molecular pathway for natural divergent evolution of function in the superfamily is unknown. We have identified single-site mutants of the (beta/alpha)(8)-barrel domains in both the l-Ala-d/l-Glu epimerase from Escherichia coli (AEE) and the muconate lactonizing enzyme II from Pseudomonas sp. P51 (MLE II) that catalyze the o-succinylbenzoate synthase (OSBS) reaction as well as the wild-type reaction. These enzymes are members of the MLE subgroup of the superfamily, share conserved lysines on opposite sides of their active sites, but catalyze acid- and base-mediated reactions with different mechanisms. A comparison of the structures of AEE and the OSBS from E. coli was used to design the D297G mutant of AEE; the E323G mutant of MLE II was isolated from directed evolution experiments. Although neither wild-type enzyme catalyzes the OSBS reaction, both mutants complement an E. coli OSBS auxotroph and have measurable levels of OSBS activity. The analogous mutations in the D297G mutant of AEE and the E323G mutant of MLE II are each located at the end of the eighth beta-strand of the (beta/alpha)(8)-barrel and alter the ability of AEE and MLE II to bind the substrate of the OSBS reaction. The substitutions relax the substrate specificity, thereby allowing catalysis of the mechanistically diverse OSBS reaction with the assistance of the active site lysines. The generation of functionally promiscuous and mechanistically diverse enzymes via single-amino acid substitutions likely mimics the natural divergent evolution of enzymatic activities and also highlights the utility of the (beta/alpha)(8)-barrel as a scaffold for new function.  相似文献   

18.
The essential catalytic base at the active site of the glycolytic enzyme triosephosphate isomerase is the carboxylate group of Glu-165, which directly abstracts either the 1-pro-R proton of dihydroxyacetone phosphate or the 2-proton of (R)-glyceraldehyde 3-phosphate to yield the cis-enediol intermediate. Using the methods of site-directed mutagenesis, we have replaced Glu-165 by Asp. The three enzymes chicken isomerase from chicken muscle, wild-type chicken isomerase expressed in Escherichia coli, and mutant (Glu-165 to Asp) chicken isomerase expressed in E. coli have each been purified to homogeneity. The specific catalytic activities of the two wild-type isomerases are identical, while the specific activity of the mutant enzyme is reduced by a factor of about 1000. The observed kinetic differences do not derive from a change in mechanism in which the aspartate of the mutant enzyme acts as a general base through an intervening water molecule, because the D2O solvent isotope effects and the stoichiometries of inactivation with bromohydroxyacetone phosphate are identical for the wild-type and mutant enzymes. Using the range of isotopic experiments that were used to delineate the free-energy profile of the wild-type chicken enzyme, we here derive the complete energetics of the reaction catalyzed by the mutant protein. Comparison of the reaction energetics for the wild-type and mutant isomerases shows that only the free energies of the transition states for the two enolization steps have been seriously affected. Each of the proton abstraction steps is about 1000-fold slower in the mutant enzyme. Evidently, the excision of a methylene group from the side chain of the essential glutamate has little effect on the free energies of the intermediate states but dramatically reduces the stabilities of the transition states for the chemical steps in the catalyzed reaction.  相似文献   

19.
D J Murphy  S J Benkovic 《Biochemistry》1989,28(7):3025-3031
The strictly conserved residue leucine-54 of Escherichia coli dihydrofolate reductase forms part of the hydrophobic wall which binds the p-aminobenzoyl side chain of dihydrofolate. In addition to the previously reported glycine-54 mutant, isoleucine-54 and asparagine-54 substitutions have been constructed and characterized with regard to their effects on binding and catalysis. NADP+ and NADPH binding is virtually unaffected with the exception of a 15-fold decrease in NADPH dissociation from the Gly-54 mutant. The synergistic effect of NADPH on tetrahydrofolate dissociation seen in the wild-type enzyme is lost in the isoleucine-54 mutant: little acceleration is seen in tetrahydrofolate dissociation when cofactor is bound, and there is no discrimination between reduced and oxidized cofactor. The dissociation constants for dihydrofolate and methotrexate increase in the order Leu less than Ile less than Asn less than Gly, varying by a maximum factor of 1700 for dihydrofolate and 6300 for methotrexate. Despite these large changes in binding affinity, the hydride transfer rate of 950 s-1 in the wild-type enzyme is decreased by a constant factor of ca. 30 (2 kcal/mol) regardless of the mutant. Thus, the contributions of residue 54 to binding and catalysis appear to have been separated.  相似文献   

20.
In this study we examine for the first time the roles of the various domains of human RNase H1 by site-directed mutagenesis. The carboxyl terminus of human RNase H1 is highly conserved with Escherichia coli RNase H1 and contains the amino acid residues of the putative catalytic site and basic substrate-binding domain of the E. coli RNase enzyme. The amino terminus of human RNase H1 contains a structure consistent with a double-strand RNA (dsRNA) binding motif that is separated from the conserved E. coli RNase H1 region by a 62-amino acid sequence. These studies showed that although the conserved amino acid residues of the putative catalytic site and basic substrate-binding domain are required for RNase H activity, deletion of either the catalytic site or the basic substrate-binding domain did not ablate binding to the heteroduplex substrate. Deletion of the region between the dsRNA-binding domain and the conserved E. coli RNase H1 domain resulted in a significant loss in the RNase H activity. Furthermore, the binding affinity of this deletion mutant for the heteroduplex substrate was approximately 2-fold tighter than the wild-type enzyme suggesting that this central 62-amino acid region does not contribute to the binding affinity of the enzyme for the substrate. The dsRNA-binding domain was not required for RNase H activity, as the dsRNA-deletion mutants exhibited catalytic rates approximately 2-fold faster than the rate observed for wild-type enzyme. Comparison of the dissociation constant of human RNase H1 and the dsRNA-deletion mutant for the heteroduplex substrate indicates that the deletion of this region resulted in a 5-fold loss in binding affinity. Finally, comparison of the cleavage patterns exhibited by the mutant proteins with the cleavage pattern for the wild-type enzyme indicates that the dsRNA-binding domain is responsible for the observed strong positional preference for cleavage exhibited by human RNase H1.  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