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 78 毫秒
1.
Parental strains and asporogenous mutants of Bacillus thuringiensis subspp. kurstaki and aizawai produced high yields of δ-endotoxin on M medium, which contained 330 μg of potassium per ml, but not on ST and ST-a media, each of which contained only 11 μg of potassium per ml. On ST and ST-a media, refractile granules were formed instead. These granules had no insecticidal activity against silkworms and were isolated and identified as poly-β-hydroxybutyric acid. Supplementation of the potassium-deficient ST-a medium with 0.1% KH2PO4 (3.7 mM) led to the formation of crystalline δ-endotoxin. The replacement of KH2PO4 with equimolar amounts of KCl, KNO3, and potassium acetate or an equivalent amount of K2SO4 had a similar effect, whereas the addition of an equimolar amount of NaH2PO4 or NH4H2PO4 did not cause the endotoxin to form. An asporogenous mutant, B. thuringiensis subsp. kurstaki strain 290-1, produced δ-endotoxin on ST-a medium supplemented with 3 mM or more potassium but formed only poly-β-hydroxybutyric acid granules on the media containing ≤1 mM potassium. These results clearly indicate that a certain concentration of potassium is essential for the fermentative production of δ-endotoxin by these isolates of B. thuringiensis. Manganese could not be substituted for potassium. Phosphate ions stimulated poly-β-hydroxybutyric acid formation by strain 290-1. The sporulation of B. thuringiensis and several other Bacillus strains was suppressed on the potassium-deficient ST medium. This suggests that potassium plays an essential role not only in Bacillus cell growth and δ-endotoxin formation but also in sporulation.  相似文献   

2.
The shuttle vector pHT3101 and its derivative pHT408, bearing a copy of a cryIA(a) δ-endotoxin gene, were transferred into several Bacillus thuringiensis subspecies through phage CP-54Ber-mediated transduction, with frequencies ranging from 5 × 10-8 to 2 × 10-6 transductant per CFU, depending on the strain and on the plasmid. In Cry- and Cry+ native recipients, the introduction of the cryIA(a) gene resulted in the formation of large bipyramidal crystals that were active against the insect Plutella xylostella (order Lepidoptera). In both cases, high levels of gene expression were observed. Transductants displaying a dual specificity were constructed by using as recipients the new isolates LM63 and LM79, which have larvicidal activity against insects of the order Coleoptera. It was not possible, however, to introduce pHT7911 into B. thuringiensis subsp. entomocidus, aizawai, or israelensis by transduction. However, electrotransformation was successful, and transformants expressing the toxin gene cryIIIA, carried by pHT7911, were obtained. Again, high levels of expression of the cloned gene were observed. The results indicate that CP-54Ber-mediated transduction is a useful procedure for introducing cloned crystal protein genes into various B. thuringiensis recipients and thereby creating strains with new combinations of genes. Finally it was also shown that pHT3101 is a very good expression vector for the cloned δ-endotoxin genes in the different recipients.  相似文献   

3.
A β-phosphoglucomutase (β-PGM) mutant of Lactococcus lactis subsp. lactis ATCC 19435 was constructed using a minimal integration vector and double-crossover recombination. The mutant and the wild-type strain were grown under controlled conditions with different sugars to elucidate the role of β-PGM in carbohydrate catabolism and anabolism. The mutation did not significantly affect growth, product formation, or cell composition when glucose or lactose was used as the carbon source. With maltose or trehalose as the carbon source the wild-type strain had a maximum specific growth rate of 0.5 h−1, while the deletion of β-PGM resulted in a maximum specific growth rate of 0.05 h−1 on maltose and no growth at all on trehalose. Growth of the mutant strain on maltose resulted in smaller amounts of lactate but more formate, acetate, and ethanol, and approximately 1/10 of the maltose was found as β-glucose 1-phosphate in the medium. Furthermore, the β-PGM mutant cells grown on maltose were considerably larger and accumulated polysaccharides which consisted of α-1,4-bound glucose units. When the cells were grown at a low dilution rate in a glucose and maltose mixture, the wild-type strain exhibited a higher carbohydrate content than when grown at higher growth rates, but still this content was lower than that in the β-PGM mutant. In addition, significant differences in the initial metabolism of maltose and trehalose were found, and cell extracts did not digest free trehalose but only trehalose 6-phosphate, which yielded β-glucose 1-phosphate and glucose 6-phosphate. This demonstrates the presence of a novel enzymatic pathway for trehalose different from that of maltose metabolism in L. lactis.  相似文献   

4.
Molasses is widely used as a substrate for commercial yeast production. The complete hydrolysis of raffinose, which is present in beet molasses, by Saccharomyces strains requires the secretion of α-galactosidase, in addition to the secretion of invertase. Raffinose is not completely utilized by commercially available yeast strains used for baking, which are Mel. In this study we integrated the yeast MEL1 gene, which codes for α-galactosidase, into a commercial mel0 baker's yeast strain. The Mel+ phenotype of the new strain was stable. The MEL1 gene was expressed when the new Mel+ baker's yeast was grown in molasses medium under conditions similar to those used for baker's yeast production at commercial factories. The α-galactosidase produced by this novel baker's yeast strain hydrolyzed all the melibiose that normally accumulates in the growth medium. As a consequence, additional carbohydrate was available to the yeasts for growth. The new strain also produced considerably more α-galactosidase than did a wild-type Mel+ strain and may prove useful for commercial production of α-galactosidase.  相似文献   

5.
Five strains of Bacillus thuringiensis that produce crystalline δ-endotoxin were used as parental strains in an effort to isolate acrystalliferous (Cry) mutants: HD-2 (B. thuringiensis var. thuringiensis, flagellar serotype 1); HD-1 and HD-73 (both var. kurstaki, serotype 3ab); HD-4 (var. alesti, serotype 3a); and HD-8 (var. galleriae, serotype 5ab). The parental strains contain complex plasmid arrays that have been previously characterized (González and Carlton, 1980). The plasmid patterns of both Cry and Cry+ variants were analyzed and compared to the parental strains using a modified Eckhardt (1978) lysate-electrophoresis method. Most Cry mutants derived from strain HD-2 were found to exhibit a distinctive colony morphology which facilitated their isolation. Loss of crystal production was associated with loss of a 75-Md plasmid. A 50-Md plasmid of strain HD-73 was lost in the Cry mutants. Crystal production in strain HD-4 appears to be associated with a plasmid about 105 Md in size; in strain HD-1, a smaller plasmid (29 Md in size) seems to be involved. In strain HD-8, a large plasmid (˜130 Md in size) is implicated in crystal production. Direct bioassay of several of the mutant strains has confirmed the loss of δ-endotoxin activity in the acrystalliferous isolates. The evidence obtained supports the notion of a relationship between specific extrachromosomal DNA elements and δ-endotoxin production in B. thuringiensis, and suggests that in each strain only a single plasmid is involved, although the size of the implicated plasmid varies from one strain to another.  相似文献   

6.
Benzo[a]pyrene is an important environmental mutagen and carcinogen. Its metabolism in cells yields the mutagenic, key ultimate carcinogen 7R,8S,9S,10R-anti-benzo[a]pyrene-7,8-dihydrodiol-9,10-epoxide, (+)-anti-BPDE, which reacts via its 10-position with N2-dG in DNA to form the adduct (+)-trans-anti-BPDE-N2-dG. To gain molecular insights into BPDE-induced mutagenesis, we examined in vivo translesion synthesis and mutagenesis in yeast cells of a site-specific 10S (+)-trans-anti-BPDE-N2-dG adduct and the stereoisomeric 10R (−)-trans-anti-BPDE-N2-dG adduct. In wild-type cells, bypass products consisted of 76% C, 14% A and 7% G insertions opposite (+)-trans-anti-BPDE-N2-dG; and 89% C, 4% A and 4% G insertions opposite (−)-trans-anti-BPDE-N2-dG. Translesion synthesis was reduced by ~26–37% in rad30 mutant cells lacking Polη, but more deficient in rev1 and almost totally deficient in rev3 (lacking Polζ) mutants. C insertion opposite the lesion was reduced by ~24–33% in rad30 mutant cells, further reduced in rev1 mutant, and mostly disappeared in the rev3 mutant strain. The insertion of A was largely abolished in cells lacking either Polη, Polζ or Rev1. The insertion of G was not detected in either rev1 or rev3 mutant cells. The rad30 rev3 double mutant exhibited a similar phenotype as the single rev3 mutant with respect to translesion synthesis and mutagenesis. These results show that while the Polζ pathway is generally required for translesion synthesis and mutagenesis of the (+)- and (−)-trans-anti-BPDE-N2-dG DNA adducts, Polη, Polζ and Rev1 together are required for G→T transversion mutations, a major type of mutagenesis induced by these lesions. Based on biochemical and genetic results, we present mechanistic models of translesion synthesis of these two DNA adducts, involving both the one-polymerase one-step and two-polymerase two-step models.  相似文献   

7.
Incubation of resting cells of Sphingobium indicum B90A, Sphingobium japonicum UT26, and Sphingobium francense Sp+ showed that they were able to transform β- and δ-hexachlorocyclohexane (β- and δ-HCH, respectively), the most recalcitrant hexachlorocyclohexane isomers, to pentachlorocyclohexanols, but only resting cells of strain B90A could further transform the pentachlorocyclohexanol intermediates to the corresponding tetrachlorocyclohexanediols. Moreover, experiments with resting cells of Escherichia coli expressing the LinB proteins of strains B90A, UT26, and Sp+ indicated that LinB was responsible for these transformations. Purified LinB proteins from all three strains also effected the formation of the respective pentachlorocyclohexanols. Although the three LinB enzymes differ only marginally with respect to amino acid sequence, they showed interesting differences with respect to substrate specificity. When LinB from strain B90A was incubated with β- and δ-HCH, the pentachlorocyclohexanol products were further transformed and eventually disappeared from the incubation mixtures. In contrast, the LinB proteins from strains UT26 and Sp+ could not catalyze transformation of the pentachlorocyclohexanols, and these products accumulated in the incubation mixture. A mutant of strain Sp+ lacking linA and linB did not degrade any of the HCH isomers, including β-HCH, and complementation of this mutant by linB from strain B90A restored the ability to degrade β- and δ-HCH.  相似文献   

8.
9.
Salmonella typhimurium strain DU501, which was found to be deficient in acetohydroxy acid synthase II (AHAS II) and to possess elevated levels of transaminase B and biosynthetic threonine deaminase, required isoleucine, methionine, or pantothenate for growth. This strain accumulated α-ketobutyrate and, to a lesser extent, α-aminobutyrate. We found that α-ketobutyrate was a competitive substrate for ketopantoate hydroxymethyltransferase, the first enzyme in pantothenate biosynthesis. This competition with the normal substrate, α-ketoisovalerate, limited the supply of pantothenate, which resulted in a requirement for methionine. Evidence is presented to support the conclusion that the ambivalent requirement for either pantothenate or methionine is related to a decrease in succinyl coenzyme A, which is produced from pantothenate and which is an obligatory precursor of methionine biosynthesis. The autointoxification by endogenously produced α-ketobutyrate could be mimicked in wild-type S. typhimurium by exogenously supplied α-ketobutyrate or salicylate, a known inhibitor of pantothenate biosynthesis. The accumulation of α-ketobutyrate was initiated by the inability of the residual AHAS activity provided by AHAS I to efficiently remove the α-ketobutyrate produced by biosynthetic threonine deaminase. The accumulation of α-ketobutyrate was amplified by the action of transaminase B, which decreased the isoleucine pool by catalyzing the formation of α-keto-β-methylvalerate and aminobutyrate from isoleucine and α-ketobutyrate; this resulted in release of threonine deaminase from end product inhibition and unbridled production of α-ketobutyrate. Isoleucine satisfied the auxotrophic requirement of the AHAS II-deficient strain by curtailing the activity of threonine deaminase. Additional lines of evidence based on genetic and physiological experiments are presented to support the basis for the autointoxification of strain DU501 as well as other nonpolarigenic ilvG mutant strains.  相似文献   

10.
Two Salmonella typhimurium strains, which could be used as sources for the leucine biosynthetic intermediates α- and β-isopropylmalate were constructed by a series of P22-mediated transductions. One strain, JK527 [flr-19 leuA2010 Δ(leuD-ara)798 fol-162], accumulated and excreted α-isopropylmalate, whereas the second strain, JK553 (flr-19 leuA2010 leuB698), accumulated and excreted α- and β-isopropylmalate. The yield of α-isopropylmalate isolated from the culture medium of JK527 was more than five times the amount obtained from a comparable volume of medium in which Neurospora crassa strain FLR92-1-216 (normally used as the source for α- and β-isopropylmalate) was grown. Not only was the yield greater, but S. typhimurium strains are much easier to handle and grow to saturation much faster than N. crassa strains. The combination of the two regulatory mutations flr-19, which results in constitutive expression of the leucine operon, and leuA2010, which renders the first leucine-specific biosynthetic enzyme insensitive to feedback inhibition by leucine, generated limitations in the production of valine and pantothenic acid. The efficient, irreversible, and unregulated conversion of α-ketoisovaleric acid into α-isopropylmalate (α-isopropylmalate synthetase Km for α-ketoisovaleric acid, 6 × 10−5 M) severely restricted the amount of α-ketoisovaleric acid available for conversion into valine and pantothenic acid (ketopantoate hydroxymethyltransferase Km for α-ketoisovaleric acid, 1.1 × 10−3 M; transaminase B Km for α-ketoisovaleric acid, 2 × 10−3 M).  相似文献   

11.
Yang Y  Yin C  Li W  Xu X 《Journal of bacteriology》2008,190(5):1554-1560
Unlike Escherichia coli, the cyanobacterium Synechocystis sp. strain PCC 6803 is insensitive to chill (5°C) in the dark but rapidly losses viability when exposed to chill in the light (100 μmol photons m−2 s−1). Preconditioning at a low temperature (15°C) greatly enhances the chill-light tolerance of Synechocystis sp. strain PCC 6803. This phenomenon is called acquired chill-light tolerance (ACLT). Preconditioned wild-type cells maintained a substantially higher level of α-tocopherol after exposure to chill-light stress. Mutants unable to synthesize α-tocopherol, such as slr1736, slr1737, slr0089, and slr0090 mutants, almost completely lost ACLT. When exposed to chill without light, these mutants showed no or a slight difference from the wild type. When complemented, the slr0089 mutant regained its ACLT. Copper-regulated expression of slr0090 from PpetE controlled the level of α-tocopherol and ACLT. We conclude that α-tocopherol is essential for ACLT of Synechocystis sp. strain PCC 6803. The role of α-tocopherol in ACLT may be based largely on a nonantioxidant activity that is not possessed by other tocopherols or pathway intermediates.  相似文献   

12.
Induction of a functional subset of HIV-specific CD4+ T cells that is resistant to HIV infection could enhance immune protection and decrease the rate of HIV disease progression. CMV-specific CD4+ T cells, which are less frequently infected than HIV-specific CD4+ T cells, are a model for such an effect. To determine the mechanism of this protection, we compared the functional response of HIV gag-specific and CMV pp65-specific CD4+ T cells in individuals co-infected with CMV and HIV. We found that CMV-specific CD4+ T cells rapidly up-regulated production of MIP-1α and MIP-1β mRNA, resulting in a rapid increase in production of MIP-1α and MIP-1β after cognate antigen stimulation. Production of β-chemokines was associated with maturational phenotype and was rarely seen in HIV-specific CD4+ T cells. To test whether production of β-chemokines by CD4+ T cells lowers their susceptibility to HIV infection, we measured cell-associated Gag DNA to assess the in vivo infection history of CMV-specific CD4+ T cells. We found that CMV-specific CD4+ T cells which produced MIP-1β contained 10 times less Gag DNA than did those which failed to produce MIP-1β. These data suggest that CD4+ T cells which produce MIP-1α and MIP-1β bind these chemokines in an autocrine fashion which decreases the risk of in vivo HIV infection.  相似文献   

13.
In Saccharomyces cerevisiae, meiosis and spore formation as well as mating are controlled by mating-type genes. Diploids heterozygous for mating type (aα) can sporulate but cannot mate; homozygous aa and αα diploids can mate, but cannot sporulate. From an αα diploid parental strain, we have isolated mutants which have gained the ability to sporulate. Those mutants which continue to mate as αα cells have been designated CSP (control of sporulation). Upon sporulation, CSP mutants yield asci containing 4α spores. The mutant gene which allows αα cells to sporulate is unlinked to the mating-type locus and also acts to permit sporulation in aa diploid cells. Segregation data from crosses between mutant αα and wild-type aa diploids and vice versa indicate (for all but one mutant) that the mutation which allows constitutive sporulation (CSP) is dominant over the wild-type allele. Some of the CSP mutants are temperature-sensitive, sporulating at 32°, but not at 23°. In addition to CSP mutants, our mutagenesis and screening procedure led to the isolation of mutants which sporulate by virtue of a change in the mating-type locus itself, resulting in loss of ability to mate.  相似文献   

14.
A Thermus thermophilus selector strain for production of thermostable and thermoactive α-galactosidase was constructed. For this purpose, the native α-galactosidase gene (agaT) of T. thermophilus TH125 was inactivated to prevent background activity. In our first attempt, insertional mutagenesis of agaT by using a cassette carrying a kanamycin resistance gene led to bacterial inability to utilize melibiose (α-galactoside) and galactose as sole carbohydrate sources due to a polar effect of the insertional inactivation. A Gal+ phenotype was assumed to be essential for growth on melibiose. In a Gal background, accumulation of galactose or its metabolite derivatives produced from melibiose hydrolysis could interfere with the growth of the host strain harboring recombinant α-galactosidase. Moreover, the AgaT strain had to be Kms for establishment of the plasmids containing α-galactosidase genes and the kanamycin resistance marker. Therefore, a suitable selector strain (AgaT Gal+ Kms) was generated by applying integration mutagenesis in combination with phenotypic selection. To produce heterologous α-galactosidase in T. thermophilus, the isogenes agaA and agaB of Bacillus stearothermophilus KVE36 were cloned into an Escherichia coli-Thermus shuttle vector. The region containing the E. coli plasmid sequence (pUC-derived vector) was deleted before transformation of T. thermophilus with the recombinant plasmids. As a result, transformation efficiency and plasmid stability were improved. However, growth on minimal agar medium containing melibiose was achieved only following random selection of the clones carrying a plasmid-based mutation that had promoted a higher copy number and greater stability of the plasmid.  相似文献   

15.
We have examined the activity of the thiamin phosphate pyrophosphorylase in Arabidopsis thaliana wild type and in a mutant (th-1) which requires exogenous thiamin for growth. Mutant and wild-type plants grown in 1 × 10−7 molar thiamin were used for the examination of the production of thiamin and thiamin monophosphate (TMP) using 4-methyl-5-hydroxyethylthiazole phosphate and 2-methyl-4-amino-5-hydroxymethylpyrimidine pyrophosphate as substrates. While the wild-type strain formed both thiamin and TMP, the th-1 mutant did not. When TMP was added to the extracts, the th-1 mutant, as well as wild type, produced thiamin. Accordingly, it was concluded that the th-1 mutant was defective in the activity of TMP pyrophosphorylase. Some of the characteristics of the enzyme from the wild-type plant were examined. The optimum temperature for the reaction is 45°C, and the Km values for the substrates are 2.7 × 10−6 molar for 4-methyl-5-hydroxyethylthiazole phosphate and 1.8 × 10−6 molar for 2-methyl-4-amino-5-hydroxymethylpyrimidine pyrophosphate.  相似文献   

16.
Uptake of [methyl-3H]thymidine label from Escherichia coli 15T- into the DNA of Dictyostelium discoideum has been measured in control and [60Co]-gamma-irradiated cells of the resistant strain NC-4 (D10, colony-forming survival = 300 krad) and two sensitive daughter strains, γs-18 (D10 = 75 krad) and γs-13 (D10 = 4 krad). Nuclear (n) and mitochondrial (m) DNA were resolved by isopycnic CsCl gradients. The uptake of label into n-DNA during the immediate postirradiation period was selectively inhibited by irradiation, compared with uptake into m-DNA. For all three strains, the gamma ray dose to reduce the uptake into n-DNA to 37% of the control during the first hour after irradiation was 3 krad, whereas for uptake into m-DNA it was 75 krad. After the initial dose- and strain-dependent lag, uptake into n-DNA resumed. γs-18 showed longer lags in n-DNA synthesis and cell division than did NC-4. γs-13 resumed n-DNA synthesis and cell division after slightly shorter lags than for NC-4. The early postlag uptake into n-DNA in this strain was almost at the control rate and was accompanied by division until the cell number had nearly doubled. The rate of label uptake then declined, division stopped, and gradual cell lysis ensued. The postdelay response of γs-13 was almost independent of dose in the range of 10-100 krad. The response of γs-18 in these and earlier experiments is consistent with the view-point that it is sensitive because of a decreased rate of repair of DNA damage. However, the basis for the sensitivity of γs-13 seems to be more complex. This strain undergoes a premature but short-lived burst of n-DNA synthesis and division for what appears to be about one round of replication. Replication then ceases, even at very low doses, leading to greatly reduced probability of survival.  相似文献   

17.
The metabolic fate of citrate and pyruvate in four strains of Lactococcus lactis subsp. lactis biovar diacetylactis has been studied by means of 13C nuclear magnetic resonance, using as a substrate either [3-13C]pyruvic acid or custom-synthesized citric acid that is 13C labeled either at carbons 2 and 4 or at carbon 3. The fermentations were carried out batchwise in modified M17 broth. For the actual conversions of the 13C-labeled substrates, cells at the end of their logarithmic growth phase were used to minimize the conversion to lactic acid. A mass balance of the main citric acid metabolites was obtained; the four strains produced from 50 to 70% (on a molar basis) lactic acid from either citrate or pyruvate. The remaining 50 to 30% was converted mainly to either α-acetolactic acid (for one strain) or acetoin (for the other three strains). One of the strains produced an exceptionally high concentration of the diacetyl precursor α-acetolactic acid. Another strain (SDC6) also produced α-acetolactic acid, but this was decarboxylated to acetoin at a high rate. The 13C nuclear magnetic resonance method confirmed that the biosynthesis of α-acetolactic acid occurs via condensation of pyruvate and “active” acetaldehyde. Diacetyl was not found as a direct metabolite of citrate or pyruvate metabolism.  相似文献   

18.
Spectral analysis indicated the presence of a cytochrome cbb3 oxidase under microaerobic conditions in Azospirillum brasilense Sp7 cells. The corresponding genes (cytNOQP) were isolated by using PCR. These genes are organized in an operon, preceded by a putative anaerobox. The phenotype of an A. brasilense cytN mutant was analyzed. Under aerobic conditions, the specific growth rate during exponential phase (μe) of the A. brasilense cytN mutant was comparable to the wild-type specific growth rate (μe of approximately 0.2 h−1). In microaerobic NH4+-supplemented conditions, the low respiration of the A. brasilense cytN mutant affected its specific growth rate (μe of approximately 0.02 h−1) compared to the wild-type specific growth rate (μe of approximately 0.2 h−1). Under nitrogen-fixing conditions, both the growth rates and respiration of the wild type were significantly diminished in comparison to those under NH4+-supplemented conditions. Differences in growth rates and respiration between the wild type and the A. brasilense cytN mutant were less pronounced under these nitrogen-fixing conditions (μe of approximately 0.03 h−1 for the wild type and 0.02 h−1 for the A. brasilense cytN mutant). The nitrogen-fixing capacity of the A. brasilense cytN mutant was still approximately 80% of that determined for the wild-type strain. This leads to the conclusion that the A. brasilense cytochrome cbb3 oxidase is required under microaerobic conditions, when a high respiration rate is needed, but that under nitrogen-fixing conditions the respiration rate does not seem to be a growth-limiting factor.  相似文献   

19.
A bacterial strain, designated KSM-7358, that can use α-cedrene for growth was isolated. The strain was identified as a member of the genus Rhodococcus and catalyzed the novel allylic oxidation of α-cedrene regiospecifically to produce (R)-10-hydroxycedrene (sec-cedrenol) with a very high yield. α-Curcumene was also produced as a possible metabolite of sec-cedrenol. A possible pathway for the microbial conversion of α-cedrene to sec-cedrenol and α-curcumene is proposed.  相似文献   

20.
"Null" mutations previously isolated at the αGpdh-1 locus of Drosophila melanogaster, because of disruption of the energy-producing α-glycerophosphate cycle, severely restrict the flight ability and relative viability of affected individuals. Two "null" alleles, αGpdh-1BO-1-4, and αGpdh-1BO-1-5, when made hemizygous with a deficiency of the αGpdh-1 locus, Df(2L)GdhA, were rendered homozygous by recombination with and selective elimination of the Df(2L)GdhA chromosome. After over 25 generations, a homozygous αGpdh-1BO-1-4 stock regained the ability to fly despite the continued absence of measurable αGPDH activity. Inter se heterozygotes of three noncomplementing αGpdh-1 "null" alleles and the "adapted" αGpdh-1BO-1-4 homozygotes were examined for metabolic enzymatic activities related to the energy-producing and pyridine nucleotide-regulating functions of the α-glycerophosphate cycle in Drosophila. The enzyme functions tested included glyceraldehyde-3-phosphate dehydrogenase, cytoplasmic and soluble malate dehydrogenase, lactate dehydrogenase, mitochondrial NADH oxidation, oxidative phosphorylation, and respiratory control with the substrates α-glycerophosphate, succinate, and pyruvate. These activities in any of the mutant genotypes in early adult life were indistinguishable from those in the wild type. There was, however, a premature deterioration and atrophy of the ultrastructural integrity of flight muscle sarcosomes observed by electron microscopy in the "null" mutants. These observations were correlated with a decrease in state 3 mitochondrial oxidation with α-glycerophosphate, succinate, and pyruvate, as well as with loss of respiratory control in adults as early as 2 wk after eclosion. Such observations, which normally are seen in aged dipterans, were accompanied by premature mortality of the mutant heterozygotes. The adapted αGpdh-1BO-1-4 was identical with wild type in each of the aging characters with the single exception of lowered rates of mitochondrial oxidative phosphorylation.  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