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 46 毫秒
1.
Random mutagenesis was performed on β-agarase, AgaB, from Zobellia galactanivorans to improve its catalytic activity and thermostability. The activities of three mutants E99K, T307I and E99K–T307I were approx. 140, 190 and 200%, respectively, of wild type β-agarase (661 U/mg) at 40°C. All three mutant enzymes were stable up to 50°C and E99K–T307I had the highest thermostability. The melting temperature (T m) of E99K–T307I, determined by CD spectra, was increased by 5.2°C over that of the wild-type enzyme (54.6°C). Activities of both the wild-type and E99K–T307I enzymes, as well as their overall thermostabilities, increased in 1 mM CaCl2. The E99K–T307I enzyme was stable at 55°C with 1 mM CaCl2, reaching 260% of the activity the wild-type enzyme held at 40°C without CaCl2.  相似文献   

2.
The regulation of formation of histidyl-transfer ribonucleic acid (tRNA) synthetase was examined in strains of Salmonella typhimurium. When the first of the histidine-forming enzymes was wild type, the presence of 2-thiazolealanine in the growth medium prevented repression of histidyl-tRNA synthetase formation elicited by the addition of 1, 2, 4-triazole-3-alanine to these cultures. Conversely, thiazolealanine had no effect on repression of histidyl-tRNA synthetase formation by triazolealanine in hisG mutant strains. These data suggest a relationship between the control of histidyl-tRNA synthetase formation and the functional state of the histidine operon.  相似文献   

3.
Abrupt changes in the Arrhenius activation energy of membrane-bound enzymes have often been correlated with changes in the physical state of membrane phospholipids. Similar changes in activation energy have also been found in soluble enzymes. The possibility exists, therefore, that in some of the membrane-bound enzymes the changes might reflect intrinsic changes of the proteins independent of changes in the membrane phospholipids. This hypothesis was investigated using Drosophila mitochondria isolated from wild type and the mutant Ocdts-1. In this mutant it has been shown that succinate-cytochrome c reductase exhibits a change in Arrhenius activation energy at 18°C which is not found in the wild type (Sondergaard, L., Nielsen, N.C. and Smillie, R.M. (1975) FEBS lett. 50, 126–129). A quantitative thin-layer chromatographic analysis of mitochondrial phospholipids showed sphingomyelin to be more abundant in the wild type than in the mutant (5.2% and 4.3% of the total phospholipids, respectively). Since it was shown that the succinate-cytochrome c reductase had a lipid requirement for full activity, reciprocal rebinding experiments were done. These experiments showed that the reconstituted membranes exhibited the change in activation energy at 18°C only when the protein moiety came from mutant mitochondria, that is, the change was independent of the source of the phospholipids used.  相似文献   

4.
Penicillium chrysogenum NRRL 792 was exposed successively to gamma radiation (physical mutagen) and ethyl methansulfonate (EMS; chemical mutagen). Gamma mutant G9 produced more alkaline protease than the wild type (62.92 vs. 40.0 U/g, respectively). Subsequent mutagenesis of G9 by EMS resulted in mutant EMS-1, which produced the highest level of enzyme (120.0 U/g). Optimal conditions for alkaline protease production by this mutant fungal strain were examined. The optimized medium was supplemented with 1 % (w/w) casein and 2.5 mM MgSO4, while the optimal pH and temperature were 9, and 30 °C after 7 days of incubation. The purified mutant alkaline protease from EMS-1 was more stable than that from the wild-type, resulting in the former having a higher pH stability and thermostability. The mutant and wild enzymes were subjected to sodium dodecylsulfate-polyacrylamide gel electrophoresis. The purified mutant enzyme showed two bands with molecular weights of 40 and 65 kDa, while the molecular weight of the purified wild-type enzyme was 66 kDa. Random amplified polymorphic DNA and inter-simple sequence repeat markers were used to identify polymorphism and genetic variations between the mutant and wild-type strains.  相似文献   

5.
The DNA ligase activities of wild type and temperature-sensitive lethal cdc 17 mutants of Schizosaccharomyces pombe have been studied by measuring effects on the conversion of relaxed DNA circles containing a single nick to a closed circular form. Such assays have revealed that all cdc 17 mutants have a thermosensitive DNA ligase deficiency, that this deficiency cosegregates 2:2 with their temperature-sensitive cdc-lethality in three tetrads derived from a cross against wild type, and that genetic reversion of the temperature-sensitive cdc? phenotype is accompanied by a restoration of DNA ligase activity; all of which implies that the temperature-sensitive cdc? phenotype of cdc 17 mutants is due to a single nuclear mutation causing a DNA ligase deficiency. Both wild type and mutant enzymes have been partially purified by chromatography in heparin/agarose columns. The wild-type enzyme is completely stable in vitro at both permissive (25 °C) and restrictive (35 °C) temperatures, whereas that of two different mutants, though completely stable at 25 °C, is rapidly inactivated at 35 °C, implying that their mutations are located in the structural gene for DNA ligase.  相似文献   

6.
We investigated chilling-induced changes in ethylene levels in Arabidopsis to find plants with distinct patterns of ethylene production in the cold-related biosynthetic pathway. The sensitive mutants identified here includedchs1-2,chs4-2, andchs6-2. Among these, plants of thechs4-2 mutant produced more ethylene than did the wild type after both were transferred from 4°C or 10°C to 22°C. This mutant also showed less freezing tolerance and more electrolyte leakage than the wild-type plants. Our results suggest a relationship between ethylene biosynthesis and chilling sensitivity in the mutant To determine which of the enzymes involved in ethylene biosynthesis were induced by chilling, we tested the activities of ACC synthase and ACC oxidase in both mutant and wild-type plants, and found greater activity by ACC synthase as well as a higher ACC content in the mutants after all the plants were transferred from 10°C to 22°C. However, ACC oxidase activity did not differ between mutant and wild-type plants in response to chilling treatment Therefore, we conclude thatchs4-2 mutants produce more ethylene than do other mutants or the wild type during their recovery from chilling conditions. Furthermore, we believe that ACC synthase is the key enzyme involved in this response.  相似文献   

7.
Histidine biosynthesis in Corynebacterium glutamicum is regulated not only by feedback inhibition by the first enzyme in the pathway, but also by repression control of the synthesis of the histidine enzymes. C. glutamicum histidine genes are located and transcribed in two unlinked loci, hisEG and hisDCB-orf1-orf2-hisHA-impA-hisFI. We constructed plasmid pK18hisDPtac to replace the native hisD promoter with the tac promoter, and overexpressed phosphoribosyl-ATP-pyrophosphohydrolase, encoded by hisE, and ATP-phosphoribosyltransferase, encoded by hisG. The l-histidine titer at 0.85 g l?1 was 80 % greater in the transformed bacterium and production of byproducts, l-alanine and l-tryptophan, was significantly decreased. However, accumulation of glutamic acid increased by 58 % (2.8 g l?1). This study represents the first attempt to substitute the histidine biosynthesis pathway promoter in the chromosome with a stronger promoter to increase histidine production.  相似文献   

8.
Mutants of simian virus 40 (SV40) with base substitutions at or near the origin of replication of the viral genome have been constructed by bisulfite mutagenesis at the BglI restriction site of SV40 DNA, followed by transfection of cells with the BglI-resistant (BglIr) DNA so generated. Based on plaque morphology at different temperatures, the resulting BglIr mutants could be classified into four-groups. Class I mutants (designated ar for “altered restriction”) were indistinguishable from wild-type SV40; class II mutants (designated shp for “sharp plaque”) produced small, sharp-edged plaques; class III mutants (designated sp for “small plaque”) produced small plaques at 32 °C, 37 °C and 40 °C; and class IV mutants (designated cs for “cold sensitive”) produced small plaques at 32 °C and wild-type plaques at 37 °C and 40 °C. That the altered plaque morphology of sp and cs mutants was related to mutation at the BglI restriction site was demonstrated by co-reversion to wild-type of the plaque phenotype and BglI sensitivity. The nucleotide sequence around the original BglI site was determined in the DNA from one mutant of each class. In each case a different base-pair substitution was found, at a site outside sequences coding for SV40 proteins. When rates of replication of mutant DNAs were measured during productive infection, ar mutant DNA was synthesized at a rate comparable to that of wild-type SV40 DNA, shp mutant DNA was made at a rate exceeding that of wild-type, sp mutant DNA was synthesized at a lower rate than that of wild type. and cs mutant DNA synthesis was reduced at 32 °C, but about the same as the wild-type rate at 40 °C. These patterns of mutant DNA synthesis were unaltered in cells co-infected with mutant and wild-type virus, i.e. the defects in DNA synthesis were not trans-complementable. We conclude that the defective mutants have single base-pair changes in a cis element that determines the rate of viral DNA replication, presumably within the origin signal itself.  相似文献   

9.
The cytoplasmic leucyl-tRNA synthetases of Neurospora crassa wild type (grown at 37 degrees C) and mutant (grown at 28 degrees C) were purified approximately 1770-fold and 1440-fold respectively. Additional enzyme preparations were carried out with mutant cells grown for 24 h at 28 degrees C and transferred then to 37 degrees C for 10-70 h of growth. The mitochondrial leucyl-tRNA synthetase of the wild type was purified approximately 722-fold. The mitochondrial mutant enzyme was found only in traces. The cytoplasmic leucyl-tRNA synthetase from the mutant (grown at 37 degrees C) in vivo is subject of a proteolytic degradation. This leads to an increased pyrophosphate exchange, without altering aminoacylation. Proteolysis in vitro by trypsin or subtilisin of isolated cytoplasmic wild-type and mutant leucyl-tRNA synthetases, however, did not establish and difference in the degradation products and in their catalytic properties. Comparing the cytoplasmic wild-type and mutant enzymes (grown at 28 degrees C) via steady-state kinetics did not show significant differences between these synthetases either. The rate-determining step appears to be after the transfer of the aminoacyl group to the tRNA, e.g. a conformational change or the release of the product. Besides leucine only isoleucine is activated by the enzymes with a discrimination of approximately 1:600; however, no Ile-tRNALeu is released. Similarly these enzymes, when tested with eight ATP analogs, cannot be distinguished. For both enzymes six ATP analogs are neither substrates nor inhibitors. Two analogs are substrates with identical kinetic parameters. The mitochondrial wild-type leucyl-tRNA synthetase is different from the cytoplasmic enzyme, as particularly exhibited by aminoacylating Escherichia coli tRNALeu but not N. crassa cytoplasmic tRNALeu. The presence of traces of the analogous mitochondrial mutant enzyme could be demonstrated. Therefore, the difference between wild-type and mutant leu-5 does not rest in the catalytic properties of the cytoplasmic leucyl-tRNA synthetases. Differences in other properties of these enzymes are not excluded. In contrast the activity of the mitochondrial leucyl-tRNA synthetase of the mutant is approximately 1% of that of the wild-type enzyme.  相似文献   

10.
A method is described for the selection of Salmonella typhimurium mutants with reduced levels of hisG enzyme activity. This method is based on the fact that the hisG enzyme catalyzes the consumption of ATP in the first step of histidine biosynthesis. Normally, this reaction is closely regulated, both by feedback inhibition and by repression of the operon. However, conditions can be set up that result in the uncontrolled use of adenine in histidine biosynthesis. Cells grown under these conditions become phenotypic adenine auxotrophs. Some revertant clones that no longer require adenine contain mutations in hisG, hisE, or the his-control region. The hisG mutations are of all types (nonsense, frameshift, missense, deletion and leaky types), and they map throughout the hisG gene.  相似文献   

11.
Mutation and immobilization techniques were applied to uridine phosphorylase (UP) from Escherichia coli in order to enhance its thermal stability and hence productivity in a biocatalytic reaction. UP was evolved by iterative saturation mutagenesis. Compared to the wild type enzyme, which had a temperature optimum of 40 °C and a half-life of 9.89 h at 60 °C, the selected mutant had a temperature optimum of 60 °C and a half-life of 17.3 h at 60 °C. Self-immobilization of the native UP as a Spherezyme showed a 3.3 fold increase in thermostability while immobilized mutant enzyme showed a 4.4 fold increase in thermostability when compared to native UP. Combining UP with the purine nucleoside phosphorylase from Bacillus halodurans allows for synthesis of 5-methyluridine (a pharmaceutical intermediate) from guanosine and thymine in a one-pot transglycosylation reaction. Replacing the wild type UP with the mutant allowed for an increase in reaction temperature to 65 °C and increased the reaction productivity from 10 to 31 g l−1 h−1.  相似文献   

12.
In order to improve the optimum temperature of lipases, the Penicillum expansum lipase (PEL) gene was mutated by site-directed mutagenesis using overlap extension PCR technique. The recombinant plasmid containing mutant E83V pPIC3.5K-lip-E83V was expressed in Pichia pastoris GS115. Comparison experiments of the mutant PEL-E83V-GS and the wild-type PEL-GS showed that the optimum temperature (45°C) of the mutant was 5°C higher than that of the wild type. The thermostability of the mutant was similar to that of the wild type. The enzymatic activity of the mutant was 188 U/ml at 37°C, which was 80% that of the wild type in the same conditions. Hydrophobic interaction may be enhanced in the surface region by the hydrophilic amino acid Glu substituted with the hydrophobic amino acid Val, and may be responsible for the improvement of the optimum temperature. Translated from Microbiology, 2005, 32(1) (in Chinese)  相似文献   

13.
14.
Incubation of maize branching enzyme, mBEI and mBEII, with 100 μM diethylpyrocarbonate (DEPC) rapidly inactivated the enzymes. Treatment of the DEPC-inactivated enzymes with 100–500 mM hydroxylamine restored the enzyme activities. Spectroscopic data indicated that the inactivation of BE with DEPC was the result of histidine modification. The addition of the substrate amylose or amylopectin retarded the enzyme inactivation by DEPC, suggesting that the histidine residues are important for substrate binding. In maize BEII, conserved histidine residues are in catalytic regions 1 (His320) and 4 (His508). His320 and His508 were individually replaced by Ala via site-directed mutagenesis to probe their role in catalysis. Expression of these mutants inE. coli showed a significant decrease of the activity and the mutant enzymes hadK m values 10 times higher than the wild type. Therefore, residues His320 and His508 do play an important role in substrate binding.  相似文献   

15.
Summary A mutant of Escherichia coli was isolated that grew at a normal rate in minimal medium at 26°C, grew at a normal rate in minimal medium at 37°C only if exogenous histidine was supplied, and grew more slowly than normal at 42°C even in the presence of histidine. In very rich media the growth rate of the mutant was normal at 26°C and 30°C, but not at 37°C or 42°C. It may be described as a temperature-conditional histidine bradytroph with a decreased ceiling to its growth rate.The histidyl-tRNA synthetase of the mutant was found to be abnormal; in crude extracts the enzyme activity was less stable and had approximately a tenfold higher apparent K Mfor histidine than normal.Under many growth conditions the histidine biosynthetic enzymes in the mutant were derepressed several hundred fold compared to the wild strain, even in the presence of exogenous genous histidine. In general, the degree of derepression in the mutant was proportional to the difference in growth rate between the mutant and normal strains; this relationship, however, did not hold below 30°C or above 37°C.The properties of the mutant could be related to the properties of its histidyl-tRNA synthetase by assuming that the enzyme participates both in protein synthesis and in histidine biosynthetic enzyme regulation and that at low temperature it functions relatively more effectively in protein synthesis than in repression, while at high temperature it functions relatively more effectively in repression.Abbreviations used tRNA transfer RNA - AICAR aminoimidazole carboxamide ribose-5-phosphate  相似文献   

16.
17.
This study aimed to improve the thermostability of alkaline α-amylase from Alkalimonas amylolytica through structure-based rational design and systems engineering of its catalytic domain. Separate engineering strategies were used to increase alkaline α-amylase thermostability: (1) replace histidine residues with leucine to stabilize the least similar region in domain B, (2) change residues (glycine, proline, and glutamine) to stabilize the highly conserved α-helices in domain A, and (3) decrease the free energy of folding predicted by the PoPMuSiC program to stabilize the overall protein structure. A total of 15 single-site mutants were obtained, and four mutants — H209L, Q226V, N302W, and P477V — showed enhanced thermostability. Combinational mutations were subsequently introduced, and the best mutant was triple mutant H209L/Q226V/P477V. Its half-life at 60 °C was 3.8-fold of that of the wild type and displayed a 3.2 °C increase in melting temperature compared with that of the wild type. Interestingly, other biochemical properties of this mutant also improved: the optimum temperature increased from 50 °C to 55 °C, the optimum pH shifted from 9.5 to 10.0, the stable pH range expanded from 7.0–11.0 to 6.0–12.0, the specific activity increased by 24 %, and the catalytic efficiency (k cat/K m) increased from 1.8×104 to 3.5?×?104 l/(g min). Finally, the mechanisms responsible for the increased thermostability were analyzed through comparative analysis of structure models. The structure-based rational design and systems engineering strategies in this study may also improve the thermostability of other industrial enzymes.  相似文献   

18.
The isolation of conditional mutants with an altered copy number of the R plasmid R1drd-19 is described. Temperature-dependent as well as amber-suppressible mutants were found. These mutant plasmids have been named pKN301 and pKN303, respectively. Both types of mutations reside on the R plasmid. No difference in molecular weight could be detected by neutral sucrose gradient centrifugation for any of the mutant plasmids when compared with the wild-type plasmid. The number of copies of the plasmids was determined by measurement of the specific activity of the R plasmid-mediated β-lactamase and by measurement of covalently closed circular (CCC) DNA in alkaline sucrose gradients and dye-CsCl density gradients. Below 34 °C the temperature-dependent mutant, pKN301, had the same copy number as the wild type, while this was four times that of the wild type above 37 °C. The amber mutant pKN303 had a copy number indistinguishable from that of the wild-type plasmid in a strain containing a strong amber suppressor and a copy number about five times that of the wild-type plasmid in a strain lacking an amber suppressor. In a strain containing a temperature-sensitive amber suppressor, the amber mutant's copy number increased with the decrease in amber suppressor activity. Thus, the existence of the temperature-dependent and the amber-suppressible R-plasmid copy mutants indicates that the system that controls the replication of plasmid R1drd-19 contains an element with a negative function and that this element is a protein.  相似文献   

19.
Ribose-5-phosphate isomerase from Clostridium thermocellum converted d-psicose to d-allose, which may be useful as a pharmaceutical compound, with no by-product. The 12 active-site residues, which were obtained by molecular modeling on the basis of the solved three-dimensional structure of the enzyme, were substituted individually with Ala. Among the 12 Ala-substituted mutants, only the R132A mutant exhibited an increase in d-psicose isomerization activity. The R132E mutant showed the highest activity when the residue at position 132 was substituted with Ala, Gln, Ile, Lys, Glu, or Asp. The maximal activity of the wild-type and R132E mutant enzymes for d-psicose was observed at pH 7.5 and 80°C. The half-lives of the wild-type enzyme at 60°C, 65°C, 70°C, 75°C, and 80°C were 11, 7.0, 4.2, 1.5, and 0.6 h, respectively, whereas those of the R132E mutant enzymes were 13, 8.2, 5.1, 3.1, and 0.9 h, respectively. The specific activity and catalytic efficiency (k cat/K m) of the R132E mutant for d-psicose were 1.4- and 1.5-fold higher than those of the wild-type enzyme, respectively. When the same amount of enzyme was used, the conversion yield of d-psicose to d-allose was 32% for the R132E mutant enzyme and 25% for the wild-type enzyme after 80 min.  相似文献   

20.
Several his mutations were found to influence nitrogen fixation in Klebsiella pneumoniae: hisB, hisC, and hisD mutants had 50% of wild-type levels of nitrogenase activity when supplied with 30 μg or less histidine/ml although this concentration did not limit protein synthesis and the mutants retained a Nif+ plate phenotype. A hisA mutation had a similar but more dramatic effect. At low concentrations of histidine the hisA mutant strain had only 5% of the nitrogenase activity found at high histidine concentration or in a his+ strain, and was also Nif- on low histidine agar plates. Addition of adenine restored nitrogenase activity in the hisA but not the hisB, hisC, or hisD mutants. Low levels of intracellular ATP, a consequence of hisG enzyme activity, correlated with loss of nitrogen-fixing ability in the hisA mutant which failed to sustain nif gene expression under these conditions. Synthesis of other major cell proteins was relatively unaffected indicating that nif gene expression is selectively regulated by the energy status of the organism.  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