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.
MDH2, the nonmitochondrial isozyme of malate dehydrogenase in Saccharomyces cerevisiae, was determined to be a target of glucose-induced proteolytic degradation. Shifting a yeast culture growing with acetate to medium containing glucose as a carbon source resulted in a 25-fold increase in turnover of MDH2. A truncated form of MDH2 lacking amino acid residues 1-12 was constructed by mutagenesis of the MDH2 gene and expressed in a haploid yeast strain containing a deletion disruption of the corresponding chromosomal gene. Measurements of malate dehydrogenase specific activity and determination of growth rates with diagnostic carbon sources indicated that the truncated form of MDH2 was expressed at authentic MDH2 levels and was fully active. However, the truncated enzyme proved to be less susceptible to glucose-induced proteolysis, exhibiting a 3.75-fold reduction in turnover rate following a shift to glucose medium. Rates of loss of activity for other cellular enzymes known to be subject to glucose inactivation were similarly reduced. An extended lag in attaining wild type rates of growth on glucose measured for strains expressing the truncated MDH2 enzyme represents the first evidence of a selective advantage for the phenomenon of glucose-induced proteolysis in yeast.  相似文献   

2.
3.
The structure of the tricarboxylic acid cycle enzyme malate dehydrogenase is highly conserved in various organisms. To test the extent of functional conservation, the rat mitochondrial enzyme and the enzyme from Escherichia coli were expressed in a strain of Saccharomyces cerevisiae containing a disruption of the chromosomal MDH1 gene encoding yeast mitochondrial malate dehydrogenase. The authentic precursor form of the rat enzyme, expressed using a yeast promoter and a multicopy plasmid, was found to be efficiently targeted to yeast mitochondria and processed to a mature active form in vivo. Mitochondrial levels of the polypeptide and malate dehydrogenase activity were found to be similar to those for MDH1 in wild-type yeast cells. Efficient expression of the E. coli mdh gene was obtained with multicopy plasmids carrying gene fusions encoding either a mature form of the procaryotic enzyme or a precursor form with the amino terminal mitochondrial targeting sequence from yeast MDH1. Very low levels of mitochondrial import and processing of the precursor form were obtained in vivo and activity could be demonstrated for only the expressed precursor fusion protein. Results of in vitro import experiments suggest that the percursor form of the E. coli protein associates with yeast mitochondria but is not efficiently internalized. Respiratory rates measured for isolated yeast mitochondria containing the mammalian or procaryotic enzyme were, respectively, 83 and 62% of normal, suggesting efficient delivery of NADH to the respiratory chain. However, expression of the heterologous enzymes did not result in full complementation of growth phenotypes associated with disruption of the yeast MDH1 gene.  相似文献   

4.
The genes encoding gluconeogenic enzymes in the nonconventional yeast Yarrowia lipolytica were found to be differentially regulated. The expression of Y. lipolytica FBP1 (YlFBP1) encoding the key enzyme fructose-1,6-bisphosphatase was not repressed by glucose in contrast with the situation in other yeasts; however, this sugar markedly repressed the expression of YlPCK1, encoding phosphoenolpyruvate carboxykinase, and YlICL1, encoding isocitrate lyase. We constructed Y. lipolytica strains with two different disrupted versions of YlFBP1 and found that they grew much slower than the wild type in gluconeogenic carbon sources but that growth was not abolished as happens in most microorganisms. We attribute this growth to the existence of an alternative phosphatase with a high Km (2.3 mM) for fructose-1,6-bisphosphate. The gene YlFBP1 restored fructose-1,6-bisphosphatase activity and growth in gluconeogenic carbon sources to a Saccharomyces cerevisiae fbp1 mutant, but the introduction of the FBP1 gene from S. cerevisiae in the Ylfbp1 mutant did not produce fructose-1,6-bisphosphatase activity or growth complementation. Subcellular fractionation revealed the presence of fructose-1,6-bisphosphatase both in the cytoplasm and in the nucleus.  相似文献   

5.
6.
Yeast NAD(+)-specific isocitrate dehydrogenase (IDH) is an octameric enzyme composed of four heterodimers of regulatory IDH1 and catalytic IDH2 subunits. The crystal structure suggested that the interactions between tetramers in the octamer are restricted to defined regions in IDH1 subunits from each tetramer. Using truncation and mutagenesis, we constructed three tetrameric forms of IDH. Truncation of five residues from the amino terminus of IDH1 did not alter the octameric form of the enzyme, but this truncation with an IDH1 G15D or IDH1 D168K residue substitution produced tetrameric enzymes as assessed by sedimentation velocity ultracentrifugation. The IDH1 G15D substitution in the absence of any truncation of IDH1 was subsequently found to be sufficient for production of a tetrameric enzyme. The tetrameric forms of IDH exhibited ~50% reductions in V(max) and in cooperativity with respect to isocitrate relative to those of the wild-type enzyme, but they retained the property of allosteric activation by AMP. The truncated (-5)IDH1/IDH2 and tetrameric enzymes were much more sensitive than the wild-type enzyme to inhibition by the oxidant diamide and concomitant formation of a disulfide bond between IDH2 Cys-150 residues. Binding of ligands reduced the sensitivity of the wild-type enzyme to diamide but had no effect on inhibition of the truncated or tetrameric enzymes. These results suggest that the octameric structure of IDH has in part evolved for regulation of disulfide bond formation and activity by ensuring the proximity of the amino terminus of an IDH1 subunit of one tetramer to the IDH2 Cys-150 residues in the other tetramer.  相似文献   

7.
8.
9.
The MDH3 isozyme of Saccharomyces cerevisiae was purified from a haploid strain containing disruptions in genomic loci encoding the mitochondrial MDH1 and nonmitochondrial MDH2 isozymes. Partial amino acid sequence analysis of the purified enzyme was conducted and used to plan polymerase chain reaction techniques to clone the MDH3 gene. The isolated gene was found to encode a 343-residue polypeptide with a molecular weight of 37,200. The deduced amino acid sequence was closely related to those of MDH1 (50% residue identity) and of MDH2 (43% residue identity). The MDH3 sequence was found to contain a carboxyl-terminal SKL tripeptide, characteristic of many peroxisomal enzymes, and immunochemical analysis was used to confirm organellar localization of the MDH3 isozyme. Levels of MDH3 were determined to be elevated in cells grown with acetate as a carbon source, and under these conditions, MDH3 contributed approximately 10% of the total cellular malate dehydrogenase activity. Disruption of the chromosomal MDH3 locus produced a reduction in cellular growth rates on acetate, consistent with the presumed function of this isozyme in the glyoxylate pathway of yeast. Combined disruption of MDH1, MDH2, and MDH3 loci in a haploid strain resulted in the absence of detectable cellular malate dehydrogenase activity.  相似文献   

10.
Among highly conserved residues in eucaryotic mitochondrial malate dehydrogenases are those with roles in maintaining the interactions between identical monomeric subunits that form the dimeric enzymes. The contributions of two of these residues, Asp-43 and His-46, to structural stability and catalytic function were investigated by construction of mutant enzymes containing Asn-43 and Leu-46 substitutions using in vitro mutagenesis of the Saccharomyces cerevisiae gene (MDH1) encoding mitochondrial malate dehydrogenase. The mutant enzymes were expressed in and purified from a yeast strain containing a disruption of the chromosomal MDH1 locus. The enzyme containing the H46L substitution, as compared to the wild type enzyme, exhibits a dramatic shift in the pH profile for catalysis toward an optimum at low pH values. This shift corresponds with an increased stability of the dimeric form of the mutant enzyme, suggesting that His-46 may be the residue responsible for the previously described pH-dependent dissociation of mitochondrial malate dehydrogenase. The D43N substitution results in a mutant enzyme that is essentially inactive in in vitro assays and that tends to aggregate at pH 7.5, the optimal pH for catalysis for the dimeric wild type enzyme.  相似文献   

11.
Cytoplasmic Lys-tRNA synthetase (LysRS) from Saccharomyces cerevisiae is a dimeric enzyme made up of identical subunits of 68 kDa. By limited proteolysis, this enzyme can be converted to a truncated dimer without loss of activity. Whereas the native enzyme strongly interacts with polyanionic carriers, the modified form displays reduced binding properties. KRS1 is the structural gene for yeast cytoplasmic LysRS. It encodes a polypeptide with an amino-terminal extension composed of about 60-70 amino acid residues, compared to its prokaryotic counterpart. This segment, containing 13 lysine residues, is removed upon proteolytic treatment of the native enzyme. The aim of the present study was to probe in vivo the significance of this amino-terminal extension. We have constructed derivatives of the KRS1 gene, encoding enzymes lacking 58 or 69 amino-terminal residues and, by site-directed mutagenesis, we have changed four or eight lysine residues from the amino-terminal segment of LysRS into glutamic acids. Engineered proteins were expressed in vivo after replacement of the wild-type KRS1 allele. The mutant enzymes displayed reduced specific activities (2-100-fold). A series of carboxy-terminal deletions, encompassing 3, 10 or 15 amino acids, were introduced into the LysRS mutants with modified amino-terminal extensions. The removal of three residues led to a 2-7-fold increase in the specific activity of the mutant enzymes. This partial compensatory effect suggests that interactions between the two extreme regions of yeast LysRS are required for a proper conformation of the native enzyme. All KRS1 derivatives were able to sustain growth of yeast cells, although the mutant cell lines displaying a low LysRS activity grew more slowly. The expression, as single-copy genes, of mutant enzymes with a complete deletion of the amino-terminal extension or with four Lys----Glu mutations, that displayed specific activities close to that of the wild-type LysRS, had no discernable effect on cell growth. We conclude that the polycationic extensions of eukaryotic aminoacyl-tRNA synthetases are dispensable, in vivo, for aminoacylation activities. The results are discussed in relation to the triggering role in in situ compartmentalization of protein synthesis that has been ascribed to the polypeptide-chain extensions that characterize most, if not all, eukaryotic aminoacyl-tRNA synthetases.  相似文献   

12.
13.
M. T. McCammon 《Genetics》1996,144(1):57-69
The two carbon compounds, ethanol and acetate, can be oxidatively metabolized as well as assimilated into carbohydrate in the yeast Saccharomyces cerevisiae. The distribution of acetate metabolic enzymes among several cellular compartments, mitochondria, peroxisomes, and cytoplasm makes it an intriguing system to study complex metabolic interactions. To investigate the complex process of carbon catabolism and assimilation, mutants unable to grow on acetate were isolated. One hundred five Acn(-) (``ACetate Nonutilizing') mutants were sorted into 21 complementation groups with an additional 20 single mutants. Five of the groups have defects in TCA cycle enzymes: MDH1, CIT1, ACO1, IDH1, and IDH2. A defect in RTG2, involved in the retrograde communication between the mitochondrion and the nucleus, was also identified. Four genes encode enzymes of the glyoxylate cycle and gluconeogenesis: ICL1, MLS1, MDH2, and PCK1. Five other genes appear to be defective in regulating metabolic activity since elevated levels of enzymes in several metabolic pathways, including the glyoxylate cycle, gluconeogenesis, and acetyl-CoA metabolism, were detected in these mutants: ACN8, ACN9, ACN17, ACN18, and ACN42. In summary, this analysis has identified at least 22 and as many as 41 different genes involved in acetate metabolism.  相似文献   

14.
15.
16.
A cDNA library from RNA of senescing cucumber cotyledons was screened for sequences also expressed in cotyledons during post-germinative growth. One clone encodes ATP-dependent phosphoenolpyruvate carboxykinase (PCK; EC 4.1.1.49), an enzyme of the gluconeogenic pathway. The sequence of a fulllength cDNA predicts a polypeptide of 74397 Da which is 43%, 49% and 57% identical to bacterial, trypanosome and yeast enzymes, respectively. The cDNA was expressed in Escherichia coli and antibodies raised against the resultant protein. The antibody recognises a single polypeptide of ca. 74 kDa, in extracts of cotyledons, leaves and roots. The cucumber genome contains a single pck gene. In the seven-day period after seed imbibition, PCK mRNA and protein steady-state levels increase in amount in cotyledons, peaking at days 2 and 3 respectively, and then decrease. Both accumulate again to a low level in senescing cotyledons. This pattern of gene expression is similar to that of isocitrate lyase (ICL) and malate synthase (MS). When green cotyledons are detached from seedlings and incubated in the dark, ICL and MS mRNAs increase rapidly in amount but PCK mRNA does not. Therefore it seems unlikely that the glyoxylate cycle serves primarily a gluconeogenic role in starved (detached) cotyledons, in contrast to post-germinative and senescing cotyledons where PCK, ICL and MS are coordinately synthesised. While exogenous sucrose greatly represses expression of icl and ms genes in dark-incubated cotyledons, it has a smaller effect on the level of PCK mRNA.  相似文献   

17.
Phospholipase C-beta 3 (PLC beta 3) is an important effector enzyme in G protein-coupled signaling pathways. Activation of PLC beta 3 by G alpha and G beta gamma subunits has been fairly well characterized, but little is known about other protein interactions that may also regulate PLC beta 3 function. A yeast two-hybrid screen of a mouse brain cDNA library with the amino terminus of PLC beta 3 has yielded potential PLC beta 3 interacting proteins including calmodulin (CaM). Physical interaction between CaM and PLC beta 3 is supported by a positive secondary screen in yeast and the identification of a CaM binding site in the amino terminus of PLC beta 3. Co-precipitation of in vitro translated and transcribed amino- and carboxyl-terminal PLC beta 3 revealed CaM binding at a putative amino-terminal binding site. Direct physical interaction of PLC beta 3 and PLC beta 1 isoforms with CaM is supported by pull-down of both isoenzymes with CaM-Sepharose beads from 1321N1 cell lysates. CaM inhibitors reduced M1-muscarinic receptor stimulation of inositol phospholipid hydrolysis in 1321N1 astrocytoma cells consistent with a physiologic role for CaM in modulation of PLC beta activity. There was no effect of CaM kinase II inhibitors, KN-93 and KN-62, on M1-muscarinic receptor stimulation of inositol phosphate hydrolysis, consistent with a direct interaction between PLC beta isoforms and CaM.  相似文献   

18.
Mapping the DNA topoisomerase III binding domain of the Sgs1 DNA helicase   总被引:1,自引:0,他引:1  
Several members of the RecQ family of DNA helicases are known to interact with DNA topoisomerase III (Top3). Here we show that the Saccharomyces cerevisiae Sgs1 and Top3 proteins physically interact in cell extracts and bind directly in vitro. Sgs1 and Top3 proteins coimmunoprecipitate from cell extracts under stringent conditions, indicating that Sgs1 and Top3 are present in a stable complex. The domain of Sgs1 which interacts with Top3 was identified by expressing Sgs1 truncations in yeast. The results indicate that the NH(2)-terminal 158 amino acids of Sgs1 are sufficient for the high affinity interaction between Sgs1 and Top3. In vitro assays using purified Top3 and NH(2)-terminal Sgs1 fragments demonstrate that at least part of the interaction is through direct protein-protein interactions with these 158 amino acids. Consistent with these physical data, we find that mutant phenotypes caused by a point mutation or small deletions in the Sgs1 NH(2) terminus can be suppressed by Top3 overexpression. We conclude that Sgs1 and Top3 form a tight complex in vivo and that the first 158 amino acids of Sgs1 are necessary and sufficient for this interaction. Thus, a primary role of the Sgs1 amino terminus is to mediate the Top3 interaction.  相似文献   

19.
20.
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