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 15 毫秒
1.
Two mutant versions of Escherichia coli aspartate transcarbamylase were created by site-specific mutagenesis. Arg-234 of the 240s loop was replaced by serine in order to help deduce the function of the interactions that normally occur between Arg-234 and both Glu-50 and Gln-231 in the R state of the enzyme. The other mutation involved the replacement of Asp-271 by asparagine to further test the functional importance of the Tyr-240-Asp-271 link that has previously been proposed to stabilize the T state of the enzyme [Middleton, S. A., & Kantrowitz, E. R. (1986) Proc. Natl. Acad. Sci. U.S.A. 83, 5866-5870]. The Arg-234----Ser holoenzyme exhibits no cooperativity, a 24-fold reduction in maximal velocity, normal affinity for carbamyl phosphate, and substantially reduced affinity for aspartate and N-(phosphonoacetyl)-L-aspartate (PALA). Unlike the wild-type enzyme, the heterotropic effectors ATP and CTP are able to influence the activity of the Arg-234----Ser enzyme at saturating aspartate concentrations. The Arg-234----Ser catalytic subunit exhibits a 33-fold reduction in maximal activity, an aspartate Km of 261 mM, compared to 5.7 mM for the wild-type catalytic subunit, and only a small alteration in the Km for carbamyl phosphate. Together these results provide additional evidence that the interdomain bridging interactions between Glu-50 of the carbamyl phosphate domain and both Arg-167 and Arg-234 of the aspartate domain are necessary for the stabilization of the high-activity-high-affinity configuration of the active site of the enzyme. Furthermore, without the interdomain bridging interactions, the holoenzyme no longer exhibits homotropic cooperativity.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Uniquely among class A beta-lactamases, the RTEM-1 and RTEM-2 enzymes contain a single disulfide bond between Cys 77 and Cys 123. To study the possible role of this naturally occurring disulfide in stabilizing RTEM-1 beta-lactamase and its mutants at residue 71, this bond was removed by introducing a Cys 77----Ser mutation. Both the wild-type enzyme and the single mutant Cys 77----Ser confer the same high levels of resistance to ampicillin in vivo to Escherichia coli; at 30 degrees C the specific activity of purified Cys 77----Ser mutant is also the same as that of the wild-type enzyme. Also, neither wild-type enzyme nor the Cys 77----Ser mutant is inactivated by brief exposure to p-hydroxymercuribenzoate. However, above 40 degrees C the mutant enzyme is less stable than wild-type enzyme. After introduction of the Cys 77----Ser mutation, none of the double mutants (containing the second mutations at residue 71) confer resistance to ampicillin in vivo at 37 degrees C; proteins with Ala, Val, Leu, Ile, Met, Pro, His, Cys, and Ser at residue 71 confer low levels of resistance to ampicillin in vivo at 30 degrees C. The use of electrophoretic blots stained with antibodies against beta-lactamase to analyze the relative quantities of mutant proteins in whole-cell extracts of E. coli suggests that all 19 of the doubly mutant enzymes are proteolyzed much more readily than their singly mutant analogues (at Thr 71) that contain a disulfide bond.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Cyclin-dependent kinases (CDKs) that control cell cycle progression are regulated in many ways, including activating phosphorylation of a conserved threonine residue. This essential phosphorylation is carried out by the CDK-activating kinase (CAK). Here we examine the effects of replacing this threonine residue in human CDK2 by serine. We found that cyclin A bound equally well to wild-type CDK2 (CDK2(Thr-160)) or to the mutant CDK2 (CDK2(Ser-160)). In the absence of activating phosphorylation, CDK2(Ser-160)-cyclin A complexes were more active than wild-type CDK2(Thr-160)-cyclin A complexes. In contrast, following activating phosphorylation, CDK2(Ser-160)-cyclin A complexes were less active than phosphorylated CDK2(Thr-160)-cyclin A complexes, reflecting a much smaller effect of activating phosphorylation on CDK2(Ser-160). The kinetic parameters for phosphorylating histone H1 were similar for mutant and wild-type CDK2, ruling out a general defect in catalytic activity. Interestingly, the CDK2(Ser-160) mutant was selectively defective in phosphorylating a peptide derived from the C-terminal domain of RNA polymerase II. CDK2(Ser-160) was efficiently phosphorylated by CAKs, both human p40(MO15)(CDK7)-cyclin H and budding yeast Cak1p. In fact, the k(cat) values for phosphorylation of CDK2(Ser-160) were significantly higher than for phosphorylation of CDK2(Thr-160), indicating that CDK2(Ser-160) is actually phosphorylated more efficiently than wild-type CDK2. In contrast, dephosphorylation proceeded more slowly with CDK2(Ser-160) than with wild-type CDK2, either in HeLa cell extract or by purified PP2Cbeta. Combined with the more efficient phosphorylation of CDK2(Ser-160) by CAK, we suggest that one reason for the conservation of threonine as the site of activating phosphorylation may be to favor unphosphorylated CDKs following the degradation of cyclins.  相似文献   

4.
Casein kinase I epsilon (CKIepsilon) is a widely expressed protein kinase implicated in the regulation of diverse cellular processes including DNA replication and repair, nuclear trafficking, and circadian rhythm. CKIepsilon and the closely related CKIdelta are regulated in part through autophosphorylation of their carboxyl-terminal extensions, resulting in down-regulation of enzyme activity. Treatment of CKIepsilon with any of several serine/threonine phosphatases causes a marked increase in kinase activity that is self-limited. To identify the sites of inhibitory autophosphorylation, a series of carboxyl-terminal deletion mutants was constructed by site-directed mutagenesis. Truncations that eliminated specific phosphopeptides present in the wild-type kinase were used to guide construction of specific serine/threonine to alanine mutants. Amino acids Ser-323, Thr-325, Thr-334, Thr-337, Ser-368, Ser-405, Thr-407, and Ser-408 in the carboxyl-terminal tail of CKIepsilon were identified as probable in vivo autophosphorylation sites. A recombinant CKIepsilon protein with serine and threonine to alanine mutations eliminating these autophosphorylation sites was 8-fold more active than wild-type CKIepsilon using IkappaBalpha as a substrate. The identified autophosphorylation sites do not conform to CKI substrate motifs identified in peptide substrates.  相似文献   

5.
Two Escherichia coli pyruvate oxidase (EC 1.2.2.2) mutant genes, poxB3 and poxB4, were cloned on plasmid pBR322. The poxB3 mutant oxidase which was described previously (Y. Y. Chang and J. E. Cronan, Jr., Proc. Natl. Acad. Sci. USA 81:4348-4352, 1984) was deficient in lipid activation but retained full catalytic activity. The poxB3 mutation was located in the C-terminal half of the gene, and the nucleotide alteration has been determined by DNA sequencing of this part of the gene and by comparing the sequence with that of the wild-type strain (C. Grabau and J. E. Cronan, Jr., submitted for publication). The poxB3 oxidase mutation is the substitution of a serine residue for Pro-536. poxB4, another pyruvate oxidase mutant gene, was also deficient in lipid activation. The major difference between the poxB3 and poxB4 oxidase was in the binding of Triton detergents. The poxB4 mutation was also located in the C-terminal half of the gene, and sequence analysis has shown that only one nucleotide base was altered, which resulted in Ala-467 being converted to a threonine residue. The results of the amino acid substitutions in the mutant proteins, leading to the functional alteration of the enzyme, are discussed.  相似文献   

6.
Activation of the serine/threonine protein kinase Akt is a multistep process. We here propose that the kinase activity of Akt is regulated via autophosphorylation in trans at two putative sites (threonine 72 and serine 246) that lie in the characteristic Akt substrate motif (RXRXX(S/T)). Incubation of Akt immunoprecipitated from transfected cells with a pre-activated Akt recombinant protein and gamma-32P-labeled ATP led to marked incorporation of radioactivity in wild-type Akt but not Akt/T72A/S246A mutant. Western blot analysis using a phosphorylated Akt substrate-specific antibody of Akt immunoprecipitated from transfected cells confirmed the autophosphorylation of wild-type Akt but not Akt/T72A/S246A mutant in insulin-like growth factor-1 (IGF-1)-stimulated cells. Autophosphorylation of Akt on Thr-72 and Ser-246 appeared to require prior phosphorylation of Akt on Thr-308 and Ser-473. Compared with wild-type Akt, Akt/T72A/S246A mutant exhibited markedly reduced basal Akt kinase activity and response to cellular stimulation by insulin-like growth factor-1, and also conferred less cellular resistance to doxorubicin-induced apoptosis. The findings from these pilot studies suggest that Akt regulates its kinase activity through autophosphorylation. Further investigation of this potential novel regulatory mechanism by which Akt performs its cellular functions is warranted.  相似文献   

7.
Using Bacillus subtilis as a host and pTB524 as a vector plasmid, we cloned the thermostable alcohol dehydrogenase (ADH-T) gene (adhT) from Bacillus stearothermophilus NCA1503 and determined its nucleotide sequence. The deduced amino acid sequence (337 amino acids) was compared with the sequences of ADHs from four different origins. The amino acid residues responsible for the catalytic activity of horse liver ADH had been clarified on the basis of three-dimensional structure. Since those catalytic amino acid residues were fairly conserved in ADH-T and other ADHs, ADH-T was inferred to have basically the same proton release system as horse liver ADH. The putative proton release system of ADH-T was elucidated by introducing point mutations at the catalytic amino acid residues, Cys-38 (cysteine at position 38), Thr-40, and His-43, with site-directed mutagenesis. The mutant enzyme Thr-40-Ser (Thr-40 was replaced by serine) showed a little lower level of activity than wild-type ADH-T did. The result indicates that the OH group of serine instead of threonine can also be used for the catalytic activity. To change the pKa value of the putative system, His-43 was replaced by the more basic amino acid arginine. As a result, the optimum pH of the mutant enzyme His-43-Arg was shifted from 7.8 (wild-type enzyme) to 9.0. His-43-Arg exhibited a higher level of activity than wild-type enzyme at the optimum pH.  相似文献   

8.
Escherichia coli K-12 minicells were employed to examine polypeptides encoded by plasmids carrying wild-type and mutant Tn1 or Tn3 transposition elements. Tn1- and Tn3-containing minicells express high levels of four transposon-specified polypeptides. Three, of molecular weights 30,000, 28,000, and 25,000, are related immunologically to beta-lactamase, the enzyme responsible for ampicillin hydrolysis. A fourth polypeptide of molecular weight 19,000 is encoded by the Tn1 or Tn3 region which spans the BamHI cleavage site. Mutant transposons which no longer produce this polypeptide transpose at higher than wild-type frequencies to give aberrant transposition products (Gill et al., J. Bacteriol. 136: 742--756, 1978; Heffron et al., Proc. Natl. Acad. Sci U.S.A. 72:3632--3627, 1975). No expression could be detected from a region of the transposons extending from the inverted repeat sequence distal to the beta-lactamase gene to more than half the distance into the Tn1 or Tn3 sequence.  相似文献   

9.
The thiol RTEM-1 beta-lactamase [Sigal, Harwood & Arentzen (1982) Proc. Natl. Acad. Sci. U.S.A. 79, 7157-7160] is inactivated by 6-beta-bromopenicillanic acid with formation of a characteristic chromophore, absorbing maximally at 350 nm, which is covalently bound to the enzyme. Model studies suggest that the chromophore is that of a 6-carboxylate thiol ester of 2,3-dihydro-2,2-dimethyl-1,4-thiazine-3,6-dicarboxylate, which can arise by rearrangement of the thiol-penicilloate obtained by thiolysis of the beta-lactam of 6-beta-bromopenicillanate. Loss of activity of the enzyme is also concerted with disappearance of its single (cysteine) thiol group. These results indicate that the thiol group of this enzyme is indeed a nucleophilic catalyst in beta-lactam turnover. The thiol beta-lactamase is also inactivated by clavulanic acid with formation of a chromophore, presumably a 3-aminoacrylate thiol ester, at 308 nm. Both 6-beta-bromopenicillanate and clavulanate are hydrolysed more slowly by the thiol enzyme than by the native serine beta-lactamase, but, probably as a consequence of this, both compounds inactivate the former enzyme more efficiently. Cefoxitin, a poor substrate of the native enzyme, does not appear to interact covalently with the thiol beta-lactamase.  相似文献   

10.
The role of the serine residue found at position 237 in the carbapenemase Sme-1 has been investigated by constructing a mutant in which Ser-237 was replaced by an alanine. The S237A mutant showed a catalytic behavior against penicillins and aztreonam very similar to that of Sme-1. By contrast, S237A was characterized by a reduced catalytic efficiency against cephems, such as cephalothin and cephaloridine. In addition, the weak activity of Sme-1 against the cephamycin cefoxitin was hardly detectable with the mutant enzyme. Finally, the Ser-237-->Ala mutation resulted in a marked decrease in catalytic activity against imipenem, showing that Ser-237 contributes to the carbapenemase activity of the class A beta-lactamase Sme-1.  相似文献   

11.
The wild-type VirA protein is known to be responsive not only to phenolic compounds but also to sugars via the ChvE protein (G. A. Cangelosi, R. G. Ankenbauer, and E. W. Nester, Proc. Natl. Acad. Sci. USA 87:6708-6712, 1990, and N. Shimoda, A. Toyoda-Yamamoto, J. Nagamine, S. Usami, M. Katayama, Y. Sakagami, and Y. Machida, Proc. Natl. Acad. Sci. USA 87:6684-6688, 1990). It is shown here that the mutant VirA(Ser-44, Arg-45) protein and the chimeric VirA-Tar protein are no longer responsive to sugars and the ChvE protein. However, whereas the chimeric VirA-Tar protein was found to be locked in a highly responsive state, the VirA(Ser-44, Arg-45) mutant protein appeared to be locked in a low responsive state. This difference turned out to be important for tumorigenicity of the host strains in virulence assays on Kalanchoë daigremontiana.  相似文献   

12.
The interferon-induced RNA-dependent protein kinase PKR is found in cells in a latent state. In response to the binding of double-stranded RNA, the enzyme becomes activated and autophosphorylated on several serine and threonine residues. Consequently, it has been postulated that autophosphorylation is a prerequisite for activation of the kinase. We report the identification of PKR sites that are autophosphorylated in vitro concomitantly with activation and examine their roles in the activation of PKR. Mutation of one site, threonine 258, results in a kinase that is less efficient in autophosphorylation and in phosphorylating its substrate, the initiation factor eIF2, in vitro. The mutant kinase is also impaired in vivo, displaying reduced ability to inhibit protein synthesis in yeast and mammalian cells and to induce a slow-growth phenotype in Saccharomyces cerevisiae. Mutations at two neighboring sites, serine 242 and threonine 255, exacerbated the effect. Taken together with earlier results (S. B. Lee, S. R. Green, M. B. Mathews, and M. Esteban, Proc. Natl. Acad. Sci. USA 91:10551-10555, 1994), these data suggest that the central part of the PKR molecule, lying between its RNA-binding and catalytic domains, regulates kinase activity via autophosphorylation.  相似文献   

13.
Undecaprenyl diphosphate (UPP) synthase catalyzes the sequential cis-condensation of isopentenyl diphosphate (IPP) onto (E,E)-farnesyl diphosphate (FPP). In our previous reports on the Micrococcus luteus B-P 26 UPP synthase, we have shown that the conserved residues in the disordered region from Ser-74 to Val-85 is crucial for the binding of FPP and the catalytic function [Fujikura, K., et al. (2000) J. Biochem. (Tokyo) 128, 917-922] and the existence of a structural P-loop motif for the FPP binding site [Fujihashi, M., et al. (2001) Proc. Natl. Acad. Sci. U.S.A., 98, 4337-4342]. To elucidate the allylic substrate binding site in more detail, we prepared eight mutant enzymes and examined their kinetic behavior. The mutant with respect to the two complementarily conserved Arg residues among the structural P-loop motif, G32R-R42G, retained the activity and showed product distribution pattern exactly similar to that of the wild-type, indicating that the complementarily conserved Arg is important for maintaining the catalytic function. Substitutions of Asp-29, Arg-33, or Arg-80 with Ala resulted in a large loss of enzyme activity, suggesting that these residues are essential for catalytic function. However, the K(m) values of these mutant enzymes for Z-GGPP, which is the first intermediate during the enzymatic cis-condensations of IPP onto FPP, were only moderately different or little changed from those of the wild type. These results suggest that the binding site for the intermediate Z-GGPP having a cis double bond is different to that for the intrinsic allylic substrate, FPP, whose diphosphate moiety is recognized by the structural P-loop.  相似文献   

14.
Endothelial NOS (eNOS)-derived NO is a key factor in regulating microvascular permeability. We demonstrated previously that eNOS translocation from the plasma membrane to the cytosol is required for hyperpermeability. Herein, we tested the hypothesis that eNOS activation in the cytosol is necessary for agonist-induced hyperpermeability. To study the fundamental properties of endothelial cell monolayer permeability, we generated ECV-304 cells that stably express cDNA constructs targeting eNOS to the cytosol or plasma membrane. eNOS-transfected ECV-304 cells recapitulate the eNOS translocation and permeability properties of postcapillary venular endothelial cells (Sánchez, F. A., Rana, R., Kim, D. D., Iwahashi, T., Zheng, R., Lal, B. K., Gordon, D. M., Meininger, C. J., and Durán, W. N. (2009) Proc. Natl. Acad. Sci. U.S.A. 106, 6849-6853). We used platelet-activating factor (PAF) as a proinflammatory agonist. PAF activated eNOS by increasing phosphorylation of Ser-1177 and inducing dephosphorylation of Thr-495, increasing NO production, and elevating permeability to FITC-dextran 70 in monolayers of cells expressing wild-type and cytosolic eNOS. PAF failed to increase permeability to FITC-dextran 70 in monolayers of cells transfected with eNOS targeted to the plasma membrane. Interestingly, this occurred despite eNOS Ser-1177 phosphorylation and production of comparable amounts of NO. Our results demonstrate that the presence of eNOS in the cytosol is necessary for PAF-induced hyperpermeability. Our data provide new insights into the dynamics of eNOS and eNOS-derived NO in the process of inflammation.  相似文献   

15.
Rat liver fructose 1,6-bisphosphatase appears to be unique in that it extends 24-26 residues beyond the COOH-terminal amino acid of other mammalian fructose 1,6-bisphosphatases and this extension contains phosphorylation sites. Using as a frame of reference the 335-residue sequence of pig kidney fructose 1,6-bisphosphatase (Marcus, F., Edelstein, I., Reardon, I., and Heinrikson, R. L. (1982) Proc. Natl. Acad. Sci. U. S. A. 79, 7161-7165), the rat liver enzyme would extend to residue 361. Limited proteolysis in the COOH-terminal region of the molecule with chymotrypsin, trypsin, or both sequentially, led us to establish that the phosphorylation sites are located at Ser residues 341 and 356. The in vitro phosphorylation of purified rat liver fructose 1,6-bisphosphatase by the catalytic subunit of cyclic AMP-dependent protein kinase results in modification at both residues, although the major site of phosphorylation (61%) is at Ser-341. In contrast, rat liver fructose 1,6-bisphosphatase purified from animals that had been injected with [32P] phosphate contains most of the label (81%) at Ser-356.  相似文献   

16.
Citrobacter freundii GN346 produces a class C beta-lactamase exhibiting the substrate profile of a typical cephalosporinase. The structural and promoter regions of the cephalosporinase gene, comprising 1408 nucleotides, were completely sequenced. The amino acid sequence of the mature enzyme, comprising 361 amino acids, and its molecular mass, 39,878 Da, were determined. The active site was confirmed to be Ser-64. The amino acid sequence of the enzyme differs from that of the cephalosporinase of C. freundii OS60 by nine residues. The nucleotide sequence of the promoter region suggests a possible attenuator structure. Lys-67, one of the most conserved residues found in class A and C beta-lactamases and penicillin-binding proteins, was converted into arginine, threonine or glutamic acid through site-directed mutagenesis. The Glu-67 enzyme had lost the catalytic activity and the Thr-67 enzyme only showed a trace of activity. The Arg-67 enzyme, which retained a significant amount of the activity, was purified. The Km values of the Arg-67 enzyme for cephalothin, cephaloridine and benzylpenicillin are 13-19 times those of the wild-type enzyme; the kcat values for the three substrates are 37%, 3%, and 36% those of the wild-type enzyme, respectively.  相似文献   

17.
beta Lys-155 in the glycine-rich sequence of the beta subunit of Escherichia coli F1-ATPase has been shown to be near the gamma-phosphate moiety of ATP by affinity labeling (Ida, K., Noumi, T., Maeda, M., Fukui, T., and Futai, M. (1991) J. Biol. Chem. 266, 5424-5429). For examination of the roles of beta Lys-155 and beta Thr-156, mutants (beta Lys-155-->Ala, Ser, or Thr; beta Thr-156-->Ala, Cys, Asp, or Ser; beta Lys-155/beta Thr-156-->beta Thr-155/beta Lys-156; and beta Thr-156/beta Val-157-->beta Ala-156/beta Thr-157) were constructed, and their properties were studied extensively. The beta Ser-156 mutant was active in ATP synthesis and had approximately 1.5-fold higher membrane ATPase activity than the wild type. Other mutants were defective in ATP synthesis, had < 0.1% of the membrane ATPase activity of the wild type, and showed no ATP-dependent formation of an electrochemical proton gradient. The mutants had essentially the same amounts of F1 in their membranes as the wild type. Purified mutant enzymes (beta Ala-155, beta Ser-155, beta Ala-156, and beta Cys-156) showed low rates of multisite (< 0.02% of the wild type) and unisite (< 1.5% of the wild type) catalyses. The k1 values of the mutant enzymes for unisite catalysis were lower than that of the wild type: not detectable with the beta Ala-156 and beta Cys-156 enzymes and 10(2)-fold lower with the beta Ala-155 and beta Ser-155 enzymes. The beta Thr-156-->Ala or Cys enzyme showed an altered response to Mg2+, suggesting that beta Thr-156 may be closely related to Mg2+ binding. These results suggest that beta Lys-155 and beta Thr-156 are essential for catalysis and are possibly located in the catalytic site, although beta Thr-156 could be replaced by a serine residue.  相似文献   

18.
In previous work (Brooker, R. J., and Wilson, T. H. (1985) Proc. Natl. Acad. Sci. U. S. A. 82, 3959-3963), lactose permease mutants were isolated which possessed an enhanced recognition for maltose. In some of these mutants, the wild-type alanine residue at position 177 was changed to valine or threonine. To gain further insight into the side chain requirement at position 177 that confers maltose recognition, further substitutions of isoleucine, leucine, phenylalanine, proline, and serine have been made via site-directed mutagenesis. Permeases containing alanine or serine exhibited poor maltose recognition whereas those containing isoleucine, leucine, phenylalanine, proline, or valine showed moderate or good recognition. As far as galactosides are concerned, the Val-177, Pro-177, and Ser-177 mutants were able to transport lactose as well as, or slightly better than, the wild-type strain. The other mutants displayed moderately reduced levels of lactose transport. For example, the Phe-177 mutant, which was the most defective, showed a level of downhill transport which was approximately 20% that of the wild-type strain. In uphill transport assays, all of the position 177 mutants were markedly defective in their ability to accumulate beta-D-thiomethylgalactopyranoside against a concentration gradient. Finally, the position 177 mutants were analyzed for their ability to catalyze an H+ leak. Interestingly, even though the wild-type permease does not leak H+ across the bacterial membrane, all of the position 177 mutants were shown to transport H+ in the absence of sugars. For most of the mutants, this H+ leak was blocked by the addition of beta-D-thiodigalactoside. Overall, these results are discussed with regard to the effects of position 177 substitutions on the sugar recognition site and H+ transport.  相似文献   

19.
W Xu  E R Kantrowitz 《Biochemistry》1991,30(9):2535-2542
Carbamoyl phosphate is held in the active site of Escherichia coli aspartate transcarbamoylase by a variety of interactions with specific side chains of the enzyme. In particular, oxygens of the phosphate of carbamoyl phosphate interact with Ser-52, Thr-53 (backbone), Arg-54, Thr-55, and Arg-105 from one catalytic chain, as well as Ser-80 and Lys-84 from an adjacent chain in the same catalytic subunit. In order to define the role of Ser-52 and Ser-80 in the catalytic mechanism, two mutant versions of the enzyme were created with Ser-52 or Ser-80 replaced by alanine. The Ser-52----Ala holoenzyme exhibits a 670-fold reduction in maximal observed specific activity, and a loss of both aspartate and carbamoyl phosphate cooperativity. This mutation also causes 23-fold and 5.6-fold increases in the carbamoyl phosphate and aspartate concentrations required for half the maximal observed specific activity, respectively. Circular dichroism spectroscopy indicates that saturating carbamoyl phosphate does not induce the same conformational change in the Ser-52----Ala holoenzyme as it does for the wild-type holoenzyme. The kinetic properties of the Ser-52----Ala catalytic subunit are altered to a lesser extent than the mutant holoenzyme. The maximal observed specific activity is reduced by 89-fold, and the carbamoyl phosphate concentration at half the maximal observed velocity increases by 53-fold while the aspartate concentration at half the maximal observed velocity increases 6-fold.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
HeLA H23 cells are a mutant female human tumor cell line harboring defective hypoxanthine phosphoribosyltransferase (HPRT; IMP-pyrophosphate phosphoribosyltransferase, EC 2.4.2.8) as a result of a mutation that alters the isoelectric point of the enzyme (G. Milman, E. Lee, G. S. Changas, J. R. McLaughlin, and J. George, Jr., Proc. Natl. Acad. Sci. USA 73:4589-4592, 1976). As shown by Milman et al. and confirmed by us here, rare HAT+ revertants arise spontaneously at 1.9 X 10(-8) frequency and express both mutant and wild-type polypeptides. Thus, the H23 mutant also carries a silent wild-type HPRT allele that is activated in revertants. To test whether the silent allele was activated via hypomethylation of genomic DNA, H23 cells were treated with inhibitors of DNA methylation, and revertants were scored by HAT or azaserine selection. At an optimal dose of 5 microM 5-azacytidine, the reversion frequency was increased about 50-fold when assayed by HAT selection and over 1,000-fold when assayed by azaserine selection. HAT+ and azaserine revertants were heterozygous for HPRT, expressing both wild-type and mutant HPRT polypeptides. Like spontaneous revertants, they contained active HPRT enzyme and were genetically unstable, reverting at about 10(-4) frequency. Similar results were found after treatment with N-methyl-N'-nitro-N-nitrosoguanidine, a DNA-alkylating agent and potent inhibitor of mammalian DNA methylation. By contrast, the DNA-ethylating agent, ethyl methanesulfonate (EMS), did not increase the HAT+ reversion frequency; it did, however, increase the frequency by which H23 revertants heterozygous for HPRT reverted to 6-thioguanine resistance. Of nine EMS revertants, seven lacked HPRT activity and had a substantially reduced expression of the wild-type polypeptide. These observations support the hypothesis that DNA methylation plays an important role in human X-chromosome inactivation and that EMS can inactivate gene expression by promoting enzymatic methylation of genomic DNA as found previously for the prolactin gene in GH3 rat pituitary tumor cells (R. D. Ivarie and J. A. Morris, Proc. Natl. Acad. Sci. USA 79:2967-2970, 1982; R. D. Ivarie, J. A. Morris, and J. A. Martial, Mol. Cell. Biol. 2:179-189, 1982).  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