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.
Aerobic ribonucleotide reductase from Escherichia coli regulates its level of activity by binding of effectors to an allosteric site in R1, located to the proposed interaction area of the two proteins that comprise the class I enzyme. Activity is increased by ATP binding and decreased by dATP binding. To study the mechanism governing this regulation, we have constructed three R1 proteins with mutations at His-59 in the activity site and one R1 protein with a mutation at His-88 close to the activity site and compared their allosteric behavior to that of the wild type R1 protein. All mutant proteins retained about 70% of wild type enzymatic activity. We found that if residue His-59 was replaced with alanine or asparagine, the enzyme lost its normal response to the inhibitory effect of dATP, whereas the enzyme with a glutamine still managed to elicit a normal response. We saw a similar result if residue His-88, which is proposed to hydrogen-bond to His-59, was replaced with alanine. Nucleotide binding experiments ruled out the possibility that the effect is due to an inability of the mutant proteins to bind effector since little difference in binding constants was observed for wild type and mutant proteins. Instead, the interaction between proteins R1 and R2 was perturbed in the mutant proteins. We propose that His-59 is important in the allosteric effect triggered by dATP binding, that the conserved hydrogen bond between His-59 and His-88 is important for the communication of the allosteric effect, and that this effect is exerted on the R1/R2 interaction.  相似文献   

2.
Kashlan OB  Cooperman BS 《Biochemistry》2003,42(6):1696-1706
Reduction of NDPs by murine ribonucleotide reductase (mRR) requires catalytic (mR1) and free radical-containing (mR2) subunits and is regulated by nucleoside triphosphate allosteric effectors. Here we present the results of several studies that refine the recently presented comprehensive model for the allosteric control of mRR enzymatic activity [Kashlan, O. B., et al. (2002) Biochemistry 41, 462-474], in which nucleotide binding to the specificity site (s-site) drives formation of an active R1(2)R2(2) dimer, ATP or dATP binding to the adenine site (a-site) drives formation of a tetramer, mR1(4a), which isomerizes to an inactive form, mR1(4b), and ATP binding to the hexamerization site (h-site) drives formation of an active R1(6)R2(6) hexamer. Analysis of the a-site D57N variant of mR1, which differs from wild-type mR1 (wt-mR1) in that its RR activity is activated by both ATP and dATP, demonstrates that dATP activation of the D57N variant RR arises from a blockage in the formation of mR1(4b) from mR1(4a), and provides strong evidence that mR1(4a) forms active complexes with mR2(2). We further demonstrate that (a) differences in the effects of ATP versus dATP binding to the a-site of wt-mR1 provide specific mechanisms by which the dATP/ATP ratio in mammalian cells could modulate in vivo RR enzymatic activity, (b) the comprehensive model is valid over a range of Mg(2+) concentrations that include in vivo concentrations, and (c) equilibrium constants derived for the comprehensive model can be used to simulate the distribution of R1 among dimer, tetramer, and hexamer forms in vivo. Such simulations indicate that mR1(6) predominates over mR1(2) in the cytoplasm of normal mammalian cells, where the great majority of RR activity is located, but that mR1(2) may be important for nuclear RR activity and for RR activity in cells in which the level of ATP is depleted.  相似文献   

3.
The RecA residues Lys248 and Glu96 are closely opposed across the RecA subunit-subunit interface in some recent models of the RecA nucleoprotein filament. The K248R and E96D single mutant proteins of the Escherichia coli RecA protein each bind to DNA and form nucleoprotein filaments but do not hydrolyze ATP or dATP. A mixture of K248R and E96D single mutant proteins restores dATP hydrolysis to 25% of the wild type rate, with maximum restoration seen when the proteins are present in a 1:1 ratio. The K248R/E96D double mutant RecA protein also hydrolyzes ATP and dATP at rates up to 10-fold higher than either single mutant, although at a reduced rate compared with the wild type protein. Thus, the K248R mutation partially complements the inactive E96D mutation and vice versa. The complementation is not sufficient to allow DNA strand exchange. The K248R and E96D mutations originate from opposite sides of the subunit-subunit interface. The functional complementation suggests that Lys248 plays a significant role in ATP hydrolysis in trans across the subunit-subunit interface in the RecA nucleoprotein filament. This could be part of a mechanism for the long range coordination of hydrolytic cycles between subunits within the RecA filament.  相似文献   

4.
The enzyme sn-glycerol-1-phosphate dehydrogenase (Gro1PDH, EC 1.1.1.261) is key to the formation of the enantiomeric configuration of the glycerophosphate backbone (sn-glycerol-1-phosphate) of archaeal ether lipids. This enzyme catalyzes the reversible conversion between dihydroxyacetone phosphate and glycerol-1-phosphate. To date, no information about the active site and catalytic mechanism of this enzyme has been reported. Using the sequence and structural information for glycerol dehydrogenase, we constructed six mutants (D144N, D144A, D191N, H271A, H287A and D191N/H271A) of Gro1PDH from Aeropyrum pernix K1 and examined their characteristics to clarify the active site of this enzyme. The enzyme was found to be a zinc-dependent metalloenzyme, containing one zinc ion for every monomer protein that was essential for activity. Site-directed mutagenesis of D144 increased the activity of the enzyme. Mutants D144N and D144A exhibited low affinity for the substrates and higher activity than the wild type, but their affinity for the zinc ion was the same as that of the wild type. Mutants D191N, H271A and H287A had a low affinity for the zinc ion and a low activity compared with the wild type. The double mutation, D191N/H271A, had no enzyme activity and bound no zinc. From these results, it was clarified that residues D191, H271 and H287 participate in the catalytic activity of the enzyme by binding the zinc ion, and that D144 has an effect on substrate binding. The structure of the active site of Gro1PDH from A. pernix K1 seems to be similar to that of glycerol dehydrogenase, despite the differences in substrate specificity and biological role.  相似文献   

5.
Ribonucleotide reductase (RNR) is a key enzyme for the synthesis of the four DNA building blocks. Class Ia RNRs contain two subunits, denoted R1 (alpha) and R2 (beta). These enzymes are regulated via two nucleotide-binding allosteric sites on the R1 subunit, termed the specificity and overall activity sites. The specificity site binds ATP, dATP, dTTP, or dGTP and determines the substrate to be reduced, whereas the overall activity site binds dATP (inhibitor) or ATP. By using gas-phase electrophoretic mobility macromolecule analysis and enzyme assays, we found that the Escherichia coli class Ia RNR formed an inhibited alpha(4)beta(4) complex in the presence of dATP and an active alpha(2)beta(2) complex in the presence of ATP (main substrate: CDP), dTTP (substrate: GDP) or dGTP (substrate: ADP). The R1-R2 interaction was 30-50 times stronger in the alpha(4)beta(4) complex than in the alpha(2)beta(2) complex, which was in equilibrium with free alpha(2) and beta(2) subunits. Studies of a known E. coli R1 mutant (H59A) showed that deficient dATP inhibition correlated with reduced ability to form alpha(4)beta(4) complexes. ATP could also induce the formation of a generally inhibited alpha(4)beta(4) complex in the E. coli RNR but only when used in combination with high concentrations of the specificity site effectors, dTTP/dGTP. Both allosteric sites are therefore important for alpha(4)beta(4) formation and overall activity regulation. The E. coli RNR differs from the mammalian enzyme, which is stimulated by ATP also in combination with dGTP/dTTP and forms active and inactive alpha(6)beta(2) complexes.  相似文献   

6.
Mammalian hexokinases are believed to have evolved from a 100-kDa hexokinase which itself is a product of duplication and fusion of an ancestral gene encoding a 50-kDa glucose 6-phosphate-sensitive hexokinase. Type II hexokinase has been shown to possess two distinct functional active sites, one in each half, which functionally resemble the original 100-kDa hexokinase, whereas type I and III isozymes possess only one active site in the C-terminal halves. This study was conducted to identify which mutations caused the loss of catalytic activity in the N-terminal halves of type I and III isozymes. Arg 174 and Ser 447 in type I isozyme and Asp 244 in type III isozyme are speculated to be the cause, because they reside adjacent to the "catalytic" site and corresponding residues, Gly 174, Asp 447, and Gly 231, are conserved in the N-terminal half of type II isozyme as well as all other 50-kDa units that possess catalytic activity. Mutations G174R and D447S in the N-terminal half of type II isozyme reduced specific activity by approximately 79 and 57%, respectively. Therefore, neither mutation alone can account for the inactivation of the N-terminal active site in type I isozyme. Either mutation, G174R or D447S, had moderate effects on Michaelis constants, K(m), for glucose and ATP. Mg(2+). Intriguingly, mutation D447S introduced a novel inhibition by unchelated ATP (K(i) = 68 microM ATP, competitive vs ATP. Mg(2+)) to the N-terminal active site of type II isozyme. Mutation G231D caused instability to type II hexokinase and near complete loss of catalytic activity (95%), suggesting that mutation G231D not only hinders catalysis at the N-terminal active site but also leads to structural instability in type II hexokinase.  相似文献   

7.
Two mutant Escherichia coli RecA proteins were prepared in which the ATP active site residue, Ser240, was replaced with asparagine and lysine (these amino acids are found in the corresponding positions in other bacterial RecA proteins). The S240N mutation had no discernible effect on the ATP-dependent activities of the RecA protein, indicating that serine and asparagine are functionally interchangeable at position 240. The S240K mutation, in contrast, essentially eliminated the ability of the RecA protein to utilize ATP as a nucleotide cofactor. The [S240K]RecA protein was able to catalyze the hydrolysis of dATP, however, suggesting that the absence of the 2'-hydroxyl group reduced an inhibitory interaction with the Lys240 side chain. Interestingly, the [S240K]RecA protein was able to promote an efficient LexA cleavage reaction but exhibited no strand exchange activity when dATP was provided as the nucleotide cofactor. This apparent separation of function may be attributable to the elevated S(0.5) value for dATP for the [S240K]RecA protein (490 μM, compared to 20-30 μM for the wild type and [S240N]RecA proteins), and may reflect a differential dependence of the LexA co-protease and DNA strand exchange activities on the nucleotide cofactor-mediated stabilization of the functionally-active state of the RecA-ssDNA complex.  相似文献   

8.
Ribonucleotide reductase reduces all four ribonucleoside diphosphates to the deoxyribonucleotides required for DNA synthesis. The enzyme is composed of two nonidentical subunits, M1 and M2. The 89-kilodalton M1 subunit contains at least two allosteric sites which, by binding nucleotide effectors, regulate the catalytic activity and substrate specificity of the enzyme. We now show that in addition, protein M1 contains a substrate-binding (catalytic) site which is specifically photolabeled after UV irradiation in the presence of the natural substrate, [32P]CDP. The photolabeling of protein M1 by [32P]CDP required the presence of the second subunit, protein M2, and ATP, the positive allosteric effector for CDP reduction. The negative effectors, dATP, dGTP, and dTTP, inhibited the photolabeling of wild type protein M1. Deoxy-ATP did not inhibit the labeling of a mutant protein M1 that is resistant to feedback inhibition by dATP. In addition, hydroxyurea and 4-methyl-5-aminoisoquinoline thiosemicarbazone, two inhibitors of ribonucleotide reductase which affect protein M2, also inhibited the [32P]CDP labeling of protein M1. These data provide new insights into the role and interaction of the two ribonucleotide reductase subunits, proteins M1 and M2, and the mechanism of action of the allosteric effectors.  相似文献   

9.
Three separate classes of ribonucleotide reductases are known, each with a distinct protein structure. One common feature of all enzymes is that a single protein generates each of the four deoxyribonucleotides. Class I and III enzymes contain an allosteric substrate specificity site capable of binding effectors (ATP or various deoxyribonucleoside triphosphates) that direct enzyme specificity. Some (but not all) enzymes contain a second allosteric site that binds only ATP or dATP. Binding of dATP to this site inhibits the activity of these enzymes. X-ray crystallography has localized the two sites within the structure of the Escherichia coli class I enzyme and identified effector-binding amino acids. Here, we have studied the regulation of three class II enzymes, one from the archaebacterium Thermoplasma acidophilum and two from eubacteria (Lactobacillus leichmannii and Thermotoga maritima). Each enzyme has an allosteric site that binds ATP or various deoxyribonucleoside triphosphates and that regulates its substrate specificity according to the same rules as for class I and III enzymes. dATP does not inhibit enzyme activity, suggesting the absence of a second active allosteric site. For the L. leichmannii and T. maritima enzymes, binding experiments also indicate the presence of only one allosteric site. Their primary sequences suggest that these enzymes lack the structural requirements for a second site. In contrast, the T. acidophilum enzyme binds dATP at two separate sites, and its sequence contains putative effector-binding amino acids for a second site. The presence of a second site without apparent physiological function leads to the hypothesis that a functional site was present early during the evolution of ribonucleotide reductases, but that its function was lost from the T. acidophilum enzyme. The other two B12 enzymes lost not only the function, but also the structural basis for the site. Also a large subgroup (Ib) of class I enzymes, but none of the investigated class III enzymes, has lost this site. This is further indirect evidence that class II and I enzymes may have arisen by divergent evolution from class III enzymes.  相似文献   

10.
The Salmonella typhimurium and Escherichia coli MutS protein is one of several methyl-directed mismatch repair proteins that act together to correct replication errors. MutS is homologous to the Streptococcus pneumoniae HexA mismatch repair protein and to the Duc1 and Rep1 proteins of human and mouse. Homology between the deduced amino acid sequence of both MutS and HexA, and the type A nucleotide binding site consensus sequence, suggested that ATP binding and hydrolysis play a role in their mismatch repair functions. We found that MutS does indeed weakly hydrolyze ATP to ADP and Pi, with a Km of 6 microM and kcat of 0.26. To show that this activity is intrinsic to MutS, we made a site-directed mutation, which resulted in the invariant lysine of the nucleotide binding consensus sequence being changed to an alanine. The mutant MutS allele was unable to complement a mutS::Tn10 mutation in vivo, and was dominant over wild type when present in high copy number. The purified mutant protein had reduced ATPase activity, with the Km affected more severely than the kcat. Like the wild type MutS protein, the mutant protein is able to bind heteroduplex DNA specifically, but the mutant protein does so with a reduced affinity.  相似文献   

11.
We examined the oligosaccharide binding to Streptomyces sp. N174 chitosanase by fluorescence spectroscopy. By means of the tryptophan fluorescence quenching, the oligosaccharide binding abilities were evaluated using the three mutant enzymes (D57A, E197A, and D201A). The enzymatic activities of the mutant enzymes were 0.5%, 20.0%, and 38.5% of that of the wild type, respectively. Scatchard plot obtained for the wild type enzyme showed a biphasic profile, suggesting that the oligosaccharide binds to the chitosanase with two different binding sites (the high affinity site and the low affinity site). In contrast, Scatchard plot for E197A exhibited a monophasic profile, in which the slope of the line corresponds to that for the low affinity binding of the wild type enzyme. A monophasic profile was also obtained for D201A, but the slope of the line was similar to that of the high affinity binding. Thus, we conclude that Glu197 and Asp201 are responsible for oligosaccharide binding at the high affinity site and the low affinity site, respectively, which correspond to the (-n) subsites and the (+n) subsites (n=1, 2, and 3). The fluorescence quenching was very weak in D57A, suggesting a strong contribution of this residue to the oligosaccharide binding.  相似文献   

12.
Ribonucleotide reductase synthesizes deoxyribonucleotides, which are essential building blocks for DNA synthesis. The mammalian ribonucleotide reductase is described as an alpha(2)beta(2) complex consisting of R1 (alpha) and R2 (beta) proteins. ATP stimulates and dATP inhibits enzyme activity by binding to an allosteric site called the activity site on the R1 protein. Despite the opposite effects by ATP and dATP on enzyme activity, both nucleotides induce formation of R1 oligomers. By using a new technique termed Gas-phase Electrophoretic-Mobility Macromolecule Analysis (GEMMA), we have found that the ATP/dATP-induced R1 oligomers have a defined size (hexamers) and can interact with the R2 dimer to form an enzymatically active protein complex (alpha(6)beta(2)). The newly discovered alpha(6)beta(2) complex can either be in an active or an inhibited state depending on whether ATP or dATP is bound. Our results suggest that this protein complex is the major form of ribonucleotide reductase at physiological levels of R1-R2 protein and nucleotides.  相似文献   

13.
The nucleotide binding properties of mutants with alterations to Asp(351) and four of the other residues in the conserved phosphorylation loop, (351)DKTGTLT(357), of sarcoplasmic reticulum Ca(2+)-ATPase were investigated using an assay based on the 2', 3'-O-(2,4,6-trinitrophenyl)-8-azidoadenosine triphosphate (TNP-8N(3)-ATP) photolabeling of Lys(492) and competition with ATP. In selected cases where the competition assay showed extremely high affinity, ATP binding was also measured by a direct filtration assay. At pH 8.5 in the absence of Ca(2+), mutations removing the negative charge of Asp(351) (D351N, D351A, and D351T) produced pumps that bound MgTNP-8N(3)-ATP and MgATP with affinities 20-156-fold higher than wild type (K(D) as low as 0.006 microM), whereas the affinity of mutant D351E was comparable with wild type. Mutations K352R, K352Q, T355A, and T357A lowered the affinity for MgATP and MgTNP-8N(3)-ATP 2-1000- and 1-6-fold, respectively, and mutation L356T completely prevented photolabeling of Lys(492). In the absence of Ca(2+), mutants D351N and D351A exhibited the highest nucleotide affinities in the presence of Mg(2+) and at alkaline pH (E1 state). The affinity of mutant D351A for MgATP was extraordinarily high in the presence of Ca(2+) (K(D) = 0.001 microM), suggesting a transition state like configuration at the active site under these conditions. The mutants with reduced ATP affinity, as well as mutants D351N and D351A, exhibited reduced or zero CrATP-induced Ca(2+) occlusion due to defective CrATP binding.  相似文献   

14.
CARD12 (Ipaf/Clan) is an important regulator of caspase-1 activation. It belongs to the family of the nucleotide-binding site and leucine-rich repeat (NBS-LRR) proteins. The NBS domain of the NBS-LRR proteins contains putative ATP/GTPase-specific P-loop and Mg2+-binding site motifs. However, the nucleotide-binding properties and the function of the NBS domain are unknown. We developed a nucleotide-binding assay and investigated nucleotide binding to CARD12. We find that the NBS domain of CARD12 contains a nucleotide-binding pocket with specificity for ATP/dATP. A point mutation in the P-loop (K175R) of the NBS domain abolishes ATP/dATP binding. We further demonstrate that the nucleotide-binding site is required for CARD12-mediated caspase-1 activation. CARD12 self-association and association with procaspase-1 in transfected cells were markedly decreased by the P-loop mutation K175R. Furthermore, the P-loop mutation greatly reduced caspase-1 activation-dependent proIL-1beta processing. Thus, CARD12 function is dependent on the nucleotide-binding site. Our data provide insights into the molecular mechanisms of CARD12-mediated caspase-1 activation.  相似文献   

15.
Using an ensemble approach, we demonstrate that an oligomeric RecA species is required for the extension phase of RecA filament formation. The RecA K72R mutant protein can bind but not hydrolyze ATP or dATP. When mixed with other RecA variants, RecA K72R causes a drop in the rate of ATP hydrolysis and has been used to study disassembly of hydrolysis-proficient RecA protein filaments. RecA K72R filaments do not form in the presence of ATP but do so when dATP is provided. We demonstrate that in the presence of ATP, RecA K72R is defective for extension of RecA filaments on DNA. This defect is partially rescued when the mutant protein is mixed with sufficient levels of wild type RecA protein. Functional extension complexes form most readily when wild type RecA is in excess of RecA K72R. Thus, RecA K72R inhibits hydrolysis-proficient RecA proteins by interacting with them in solution and preventing the extension phase of filament assembly.  相似文献   

16.
Mucopolysaccharidosis IVA (MPS IVA) is caused by the deficiency of the lysosomal enzyme N-acetylgalactosamine-6-sulfate sulfatase encoded by the GALNS gene on chromosome 16. We describe, in detail, the clinical phenotype of five patients from three unrelated Finnish families and have characterized the disease-causing mutations in GALNS. Genotypes of the patients are D60N/A291T, D60N/W230X, and D60N/1374delT. Mutation 1374delT introduces premature termination of GALNS. Cells over-expressing the novel mutation W230X and A291T had no residual GALNS activity, whereas D60N gave 12.2% residual activity compared with the wild type. Co-transfection of D60N/A291T and D60N/W230X showed 5.5% and 6.7% of wild type activity, respectively. The precursor proteins of D60N and A291T were observed at 55 kDa and 57 kDa, respectively, whereas there was no detectable band in cells over-expressing W230X. At 55 degrees C, the mutant protein showed lower thermostability than the wild type protein at pH 3.8 and 7.0. The tertiary structural model of the GALNS protein revealed that aspartic acid at position 60 is located on the surface of the molecule, away from the active site. This makes it unlikely that the enzymatic function of the protein with D60N is severely impaired. On the other hand, A291 and W230 are localized near the active site. The molecular characteristics of the D60N mutation explain the attenuated clinical phenotype of the patients.  相似文献   

17.
The purified RNA polymerase complex of vesicular stomatitis virus required added thiols for maximal activity, whereas polymerase activity from whole disrupted virions did not. Maximal activity of the purified polymerase complex required greater than or equal to 1 mM added dithiothreitol. The polymerase was inactivated by N-ethylmaleimide (NEM) at 0 degree C, with k2 = 528 +/- 26 M-1 min-1. Activity was recovered by addition of L protein, but not N or NS, to the NEM-inactivated complex, indicating that the NEM-sensitive group was present on the L protein. Nucleoside triphosphates protected the enzyme against inactivation by N-ethylmaleimide. ATP was most effective, with KD = 0.58 +/- 0.07 mM, a value close to the Km of ATP reported previously for initiation of RNA synthesis. dATP was nearly as effective, and GTP was slightly less effective than ATP. Non-hydrolyzable analogs of ATP protected weakly, whereas ADP and pyrimidine triphosphates gave very poor, but still measurable, protection. The ATP binding site thus identified differs from the protein kinase-associated ATP binding site identified on L protein by Sanchez et al. (Sanchez, A., De, B.P., and Banerjee, A. K. (1985) J. Gen. Virol. 66, 1025-1036) in having a substantially lower affinity for ATP. Two putative ATP binding sites were identified in the L protein amino acid sequence, but none were found in the N or NS sequences.  相似文献   

18.
The cystic fibrosis transmembrane conductance regulator (CFTR) is an anion channel in the ATP-binding cassette (ABC) transporter family. CFTR consists of two transmembrane domains, two nucleotide-binding domains (NBD1 and NBD2), and a regulatory domain. Previous biochemical reports suggest NBD1 is a site of stable nucleotide interaction with low ATPase activity, whereas NBD2 is the site of active ATP hydrolysis. It has also been reported that NBD2 additionally possessed adenylate kinase (AK) activity. Knowledge about the intrinsic biochemical activities of the NBDs is essential to understanding the Cl(-) ion gating mechanism. We find that purified mouse NBD1, human NBD1, and human NBD2 function as adenylate kinases but not as ATPases. AK activity is strictly dependent on the addition of the adenosine monophosphate (AMP) substrate. No liberation of [(33)P]phosphate is observed from the gamma-(33)P-labeled ATP substrate in the presence or absence of AMP. AK activity is intrinsic to both human NBDs, as the Walker A box lysine mutations abolish this activity. At low protein concentration, the NBDs display an initial slower nonlinear phase in AK activity, suggesting that the activity results from homodimerization. Interestingly, the G551D gating mutation has an exaggerated nonlinear phase compared with the wild type and may indicate this mutation affects the ability of NBD1 to dimerize. hNBD1 and hNBD2 mixing experiments resulted in an 8-57-fold synergistic enhancement in AK activity suggesting heterodimer formation, which supports a common theme in ABC transporter models. A CFTR gating mechanism model based on adenylate kinase activity is proposed.  相似文献   

19.
In an attempt to distinguish simian virus 40 (SV40) large T antigen (T) binding to ATP from hydrolysis, specific mutations were made in the ATP-binding site of T according to our model for the site (M. K. Bradley, T. F. Smith, R. H. Lathrop, D. M. Livingston, and T. A. Webster, Proc. Natl. Acad. Sci. USA 84:4026-4030, 1987). Two acidic residues predicted to make contact with the magnesium phosphate were changed to alanines. The mutated T gene was completely defective for viral DNA synthesis and for virion production, and it was dominant defective for viral DNA replication. The defective T gene encoded a stable product (2905T) that oncogenically transformed mouse cell lines. 2905T, immunoprecipitated from transformed-cell extracts, bound SV40 origin DNA specifically and, surprisingly, it was active as an ATPase. A recombinant baculovirus was constructed for the production and purification of the mutant protein for detailed biochemical analyses. 2905T had only 10% of the ATPase and helicase of wild-type T. The Km of 2905T for ATP in ATPase assays was the same as the Km of wild-type T. ATP activated the ATPase activity of wild-type T, but not of 2905T. As tested by gel bandshift assay, 2905T bound to SV40 origin DNA and to individual sites I and II with affinities similar to that of the wild type. However, ATP did not modulate the DNA-binding activity of mutant T to site II. Therefore, this mutation in the ATP-binding site in T resulted in defects in the interaction between the protein and ATP that appeared to be responsible for the determination of the active state of T for DNA binding versus ATPase.  相似文献   

20.
Two mutant forms of fumarase C from E. coli have been made using PCR and recombinant DNA. The recombinant form of the protein included a histidine arm on the C-terminal facilitating purification. Based on earlier studies, two different carboxylic acid binding sites, labeled A- and B-, were observed in crystal structures of the wild type and inhibited forms of the enzyme. A histidine at each of the sites was mutated to an asparagine. H188N at the A-site resulted in a large decrease in specific activity, while the H129N mutation at the B-site had essentially no effect. From the results, we conclude that the A-site is indeed the active site, and a dual role for H188 as a potential catalytic base is proposed. Crystal structures of the two mutant proteins produced some unexpected results. Both mutations reduced the affinity for the carboxylic acids at their respective sites. The H129N mutant should be particularly useful in future kinetic studies because it sterically blocks the B-site with the carboxyamide of asparagine assuming the position of the ligand's carboxylate. In the H188N mutation at the active site, the new asparagine side chain still interacts with an active site water that appears to have moved slightly as a result of the mutation.  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