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.
2.
3.
Yeast cells deficient in DNA ligase were also deficient in their capacity to rejoin single-strand scissions in prelabeled nuclear DNA. After high-dose-rate gamma irradiation (10 and 25 krads), cdc9-9 mutant cells failed to rejoin single-strand scissions at the restrictive temperature of 37 degrees C. In contrast, parental (CDC9) cells (incubated with mutant cells both during and after irradiation) exhibited rapid medium-independent DNA rejoining after 10 min of post-irradiation incubation and slower rates of rejoining after longer incubation. Parental cells were also more resistant than mutant cells to killing by gamma irradiation. Approximately 2.5 +/- 0.07 and 5.7 +/- 0.6 single-strand breaks per 10(8) daltons were detected in DNAs from either CDC9 or cdc9-9 cells converted to spheroplasts immediately after 10 and 25 krads of irradiation, respectively. At the permissive temperature of 23 degrees C, the cdc9-9 cells contained 2 to 3 times the number of DNA single-strand breaks as parental cells after 10 min to 4 h of incubation after 10 krads of irradiation, and two- to eightfold more breaks after 10 min to 2.5 h of incubation after 25 krads of irradiation. Rejoining of single-strand scissions was faster in medium. After only 10 min in buffered growth medium and after 10 krads of irradiation, the number of DNA single-strand breaks was reduced to 0.32 +/- 0.3 (at 23 degrees C) or 0.21 +/- 0.05 (at 37 degrees C) per 10(8) daltons in parental cells, but remained at 2.1 +/- 0.06 (at 23 degrees C) or 2.3 +/- 0.07 (at 37 degrees C) per 10(8) daltons in mutant cells. After 10 or 25 krads of irradiation plus 1 h of incubation in medium at 37 degrees C, only DNA from CDC9 cells was rejoined to the size of DNA from unirradiated cells, whereas at 23 degrees C, DNAs in both strains were completely rejoined.  相似文献   

4.
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   

5.
The Caulobacter polar surface structures (flagella, pili, and the deoxyribonucleic acid phage phiCbK receptors), which are expressed at proximal sites of swarmer cells in a coordinate manner (Shapiro, Annu. Rev. Microbiol., 30:377-407, 1976) could be blocked by a single mutation. The mutant C. crescentus CB13 ple-801 did not form these surface structures when grown at 35 degrees C. Upon shift down to 25 degrees C, the mutant cells initiated the formation of the surface structures. When mitomycin C was added to the mutant culture upon shift down from 35 to 25 degrees C, phiCbK receptor formation was inhibited to a minimal level. Rifampin and chloramphenicol completely inhibited phiCbK receptor formation when added to the mutant culture upon shift down. Deoxyribonucleic acid as well as ribonucleic acid and protein synthesis seem to be required for the formation of phiCbK receptors. Penicillin V also inhibited phiCbK receptor formation, indicating the involvement of cell wall synthesis. When the mutant CB13 ple-801 cells were shifted down briefly from 35 to 25 degrees C and then shifted up to 35 degrees C, flagella and phiCbK receptors were formed even at 35 degrees C to different extents depending on how long the cells were incubated at 25 degrees C. This formation of the surface structures at 35 degrees C was inhibited by rifampin. From these results, it appears that translation, assembly, or localization processes for the formation of the surface structures are not temperature sensitive at 35 degrees C in the pleiotropic mutant CB13 ple-801. The syntheses of deoxyribonucleic acid and the cell wall do not appear to be temperature sensitive either, since the mutant grows normally at 35 degrees C. It is suggested that there exists a regulatory step that commits the cells to initiate the synthesis of requisite ribonucleic acid for the formation of the polar surface structures.  相似文献   

6.
Growth of temperature-sensitive mutant Bacillus cereus T JS22-C occurred normally at the restrictive temperature (37 degrees C), but sporulation was blocked at stage 0. The production of extracellular and intracellular proteases and of alkaline phosphatase occurred at 37 degrees C, but the expression of a functional tricarboxylic acid cycle did not. At the permissive temperature (26 degrees C), the mutant sporulated at a slightly lower frequency (60%) and at a lower rate than the parent strain. The oxidation of organic acids, which accumulate in the growth medium began at T0 in cultures of the parent strain but was delayed until about T3 in cultures of the mutant. Later events in sporulation were also delayed in the mutant by about 3 h. Experiments in which the temperature of growth was shifted from 37 to 26 degrees C or from 26 to 37 degrees C at various times showed that the temperature-sensitive event began approximately 1 h after the end of exponential growth and ended when the cells reached the end of stage II (septum formation). The absence of a functional tricarboxylic acid cycle in cells of the mutant grown at 37 degrees C or shifted from 26 to 37 degrees C before T1 did not appear to be due to a lesion in one of the structural genes of the tricarboxylic acid cycle but was more likely due to the inability of the cells to derepress the synthesis of some of the enzymes of that cycle.  相似文献   

7.
SSH1, a newly identified member of the heat shock protein (hsp70) multigene family of the budding yeast Saccharomyces cerevisiae, encodes a protein localized to the mitochondrial matrix. Deletion of the SSH1 gene results in extremely slow growth at 23 degrees C or 30 degrees C, but nearly wild-type growth at 37 degrees C. The matrix of the mitochondria contains another hsp70, Ssc1, which is essential for growth and required for translocation of proteins into mitochondria. Unlike SSC1 mutants, an SSH1 mutant showed no detectable defects in import of several proteins from the cytosol to the matrix compared to wild type. Increased expression of Ssc1 partially suppressed the cold- sensitive growth defect of the SSH1 mutant, suggesting that when present in increased amounts, Ssc1 can at least partially carry out the normal functions of Ssh1. Spontaneous suppressors of the cold-sensitive phenotype of an SSH1 null mutant were obtained at a high frequency at 23 degrees C, and were all found to be respiration deficient. 15 of 16 suppressors that were analyzed lacked mitochondrial DNA, while the 16th had reduced amounts. We suggest that Ssh1 is required for normal mitochondrial DNA replication, and that disruption of this process in ssh1 cells results in a defect in mitochondrial function at low temperatures.  相似文献   

8.
A psychrotrophic pseudomonad isolated from iced fish oxidized alanine at temperatures close to 0 degrees C and grew over the range 0 degrees C-35 degrees C. The rate of oxidation of alanine, measured manometrically, by cells grown at 2 degrees C was lower than that of cells grown at 22 degrees C. However, the consumption of oxygen after heat treatment at 35 degrees for 35 min was reduced considerably by 2 degrees C grown cells. Alanine oxidase activity was tested in an extract from cells grown at 2 degrees C and 22 degrees C with alanine as the sole carbon, nitrogen, and energy source. Cells grown at 2 degrees C produced an alanine oxidase with a temperature optimum of 35 degrees C and pH optimum of 8, which lost about 80% activity by heat treatment at 40 degrees C for 30 min. There was no change in activity after dialysis at pH 7, 8, or 9. Extracts from cells grown at 22 degrees C contained an alanine oxidase system with an optimum temperature of 45 degrees C, a pH optimum above 8, and only about 30% reduction of activity after heat treatment. This enzyme activity was concentrated in the 0.5 M elution fraction from a Sephadex column, and dialysis reduced the activity at pH 7 and 8. Mesophilic enzyme synthesis apparently started around a growth temperature of 10 degrees C. The crude alanine oxidase systems of Pseudomonas aeruginosa derived from cells grown at 13 degrees C and 37 degrees C had a common optimum temperature of 45 degrees C. These data suggest that one mechanism of psychrophilic growth by psychrotrophic bacteria may be the induction of enzymes with low optimum temperatures in response to low temperature conditions.  相似文献   

9.
We have isolated a thermosensitive mutant which is transformed into a population of cells devoid of mitochondrial DNA (rho 0 cells) at 35 degrees C and is deficient in mitochondrial (mt) DNA polymerase activity. A single recessive nuclear mutation (mip1) is responsible for rho 0 phenotype and mtDNA polymerase deficiency in vitro. At 25 degrees C (or 30 degrees C) a dominant suppressor mutation (SUP) masks the deficiency in vivo. The meiotic segregants (mip1 sup) which do not harbor the suppressor have a rho 0 phenotype both at 25 and 35 degrees C. They have no mtDNA polymerase activity, in contrast with MIP rho 0 mutants of mitochondrial inheritance which do exhibit mtDNA polymerase activity. In the thermosensitive mutant (mip1 SUP), the replication of mtDNA observed in vivo at 30 degrees C is completely abolished at 35 degrees C. In the meiotic segregants (mip1 sup), no mtDNA replication takes place at 30 and 35 degrees C. The synthesis of nuclear DNA is not affected. DNA polymerases may have replicative and/or repair activity. There is no evidence that mip mutants are deficient in mtDNA repair. In contrast the MIP gene product is strictly required for the replication of mtDNA and for the expression of the mtDNA polymerase activity. This enzyme might be the replicase of mtDNA.  相似文献   

10.
We have investigated mitochondrial import and processing of the precursor for human ornithine transcarbamylase (OTC; carbamoylphosphate:L-ornithine carbamoyltransferase, EC 2.1.3.3) in HeLa cells stably transformed with cDNA sequences encoding OTC precursors carrying mutations in their leader peptides. The mutant precursors studied included two with amino acid substitutions in the 32-amino-acid leader peptide (glycine for arginine at position 23, designated gly23; glycines for arginines at positions 15, 23, and 26, designated gly15,23,26) and two with deletions (deletion of residues 8 to 22, designated d8-22; deletion of residues 17 to 32, designated N16). Specific immunoprecipitation with anti-OTC antiserum of extracts of L-[35S]methionine-labeled cells expressing these mutations yielded only precursor species; neither mature nor intermediate-size OTC subunits were observed. Fractionation of radiolabeled cells, however, revealed important differences among the various mutants: the gly23 precursor was associated with mitochondria and was not detected in the cytosol; the d8-22 and N16 precursors were found with both the mitochondrial fraction and the cytosol; only the gly15,23,26 precursor was detected exclusively in the cytosol. A large fraction of each of the mitochondrially associated OTC species was in a trypsin-protected compartment. In particular, the gly23 precursor behaved in trypsin protection and mitochondrial fractionation studies in a manner consistent with its translocation into the mitochondrial matrix. On the other hand, the lack of binding of the gly23 protein to a delta-N-phosphonoacetyl-L-ornithine affinity column, which specifically recognizes active OTC enzyme, indicated that, despite its intramitochondrial location, the mutant protein did not assemble into the normal, active trimer. Further, the gly23 mutant precursor was unstable within the mitochondria and was degraded with a t1/2 of less further than 4 h. Thus, we have shown that, in intact HeLa cells, cleavage of the OTC leader peptide is not required for translocation into mitochondria, but is required for assembly into active enzyme.  相似文献   

11.
Utilizing a temperature sensitive p53 mutant (pLTRp53cGval135) which expresses mutant p53 at 37 degrees C and a wild-type like p53 at 32 degrees C, we transfected a human ovarian cancer cell line (SKOV3) which does not express endogenous p53. Among the different clones obtained, we selected three clones. Two were obtained from simultaneous transfection of p53 and neomycin resistance expression plasmids (SK23a and SK9), the other was obtained from transfection experiments utilizing the neomycin resistance gene only (SKN). Introduction of mutant p53 did not alter the morphology or growth characteristics of this ovarian cancer cell line. Upon shifting to the permissive temperature, a dramatic change in morphology and growth rate was observed in SK23a and SK9 cells that is associated with the presence of a wild-type like p53. SKN and SKOV3 cells maintained at 32 degrees C did not change morphology and only slightly reduced proliferation. Both SK23a and SK9 cells did not show evidence of apoptosis when measured up to 72 hours of maintenance at 32 degrees C. In contrast to what observed in other cell lines, SK23a and SK9 cells maintained at 32 degrees C were not blocked in G1, but they were accumulated in G2-M. This accumulation was transient and could be due either to a blockade or to a delay in the G2 progression. No down-regulation of c-myc was observed in p53 expressing clones when shifted to the permissive temperature. In these conditions gadd45 mRNA expression was highly stimulated in SK9 and SK23a cells but not in SKN cells. In both clones Gas1 mRNA was not detected either at 37 degrees C or 32 degrees C. This system represents a new and useful model for studying the effect of the absence of p53 (SKOV3 or SKN), presence of mutated p53 (SK23a and SK9 kept at 37 degrees C) or wild type p53 (SK23a and SK9 kept at 32 degrees C) on the mechanism of response of cancer cells to DNA damaging agents.  相似文献   

12.
We have identified two temperature-sensitive peroxisome-deficient mutants of Hansenula polymorpha (ts6 and ts44) within a collection of ts mutants which are impaired for growth on methanol at 43 degrees C but grow well at 35 degrees C. In both strains peroxisomes were completely absent in cells grown at 43 degrees C; the major peroxisomal matrix enzymes alcohol oxidase, dihydroxyacetone synthase and catalase were synthesized normally but assembled into the active enzyme protein in the cytosol. As in wild-type cells, these enzymes were present in peroxisomes under permissive growth conditions (< or = 37 degrees C). However, at intermediate temperatures (38-42 degrees C) they were partly peroxisome-bound and partly resided in the cytosol. Genetic analysis revealed that both mutant phenotypes were due to monogenic recessive mutations mapped in the same gene, designated PER13. After a shift of per13-6ts cells from restrictive to permissive temperature, new peroxisomes were formed within 1 h. Initially one--or infrequently a few--small organelles developed which subsequently increased in size and multiplied by fission during prolonged permissive growth. Neither mature peroxisomal matrix nor membrane proteins, which were present in the cytosol prior to the temperature shift, were incorporated into the newly formed organelles. Instead, these proteins remained unaffected (and active) in the cytosol concomitant with further peroxisome development. Thus in H.polymorpha alternative mechanisms of peroxisome biogenesis may be possible in addition to multiplication by fission upon induction of the organelles by certain growth substrates.  相似文献   

13.
The nuclear mas5 mutation causes temperature-sensitive growth and defects in mitochondrial protein import at the nonpermissive temperature in the yeast Saccharomyces cerevisiae. The MAS5 gene was isolated by complementation of the mutant phenotypes, and integrative transformation demonstrated that the complementing fragment encoded the authentic MAS5 gene. The deduced protein sequence of the cloned gene revealed a polypeptide of 410 amino acids which is homologous to Escherichia coli DnaJ and the yeast DnaJ log SCJ1. Northern (RNA blot) analysis revealed that MAS5 is a heat shock gene whose expression increases moderately at elevated temperatures. Cells with a deletion mutation in MAS5 grew slowly at 23 degrees C and were inviable at 37 degrees C, demonstrating that MAS5 is essential for growth at increased temperatures. The deletion mutant also displayed a modest import defect at 23 degrees C and a substantial import defect at 37 degrees C. These results indicate a role for a DnaJ cognate protein in mitochondrial protein import.  相似文献   

14.
Host dna functions involved in the replication of microvirid phage phiC DNA were investigated in vivo. Although growth of this phage was markedly inhibited even at 35-37 degrees C even in dna+ host, conversion of the infecting single-stranded DNA into the double-stranded parental replicative form (stage I synthesis) occurred normally at 43 degrees C in dna+, dnaA, dnaB, dnaC(D), and dnaE cells. In dnaG mutant, the stage I synthesis was severely inhibited at 43 degrees C but not at 30 degrees C. The stage I replication of phiC DNA was clearly thermosensitive in dnaZ cells incubated in nutrient broth. In Tris-casamino acids-glucose medium, however, the dnaZ mutant sufficiently supported synthesis of the parental replicative form. At 43 degrees C, synthesis of the progeny replicative form DNA (stage II replication) was significantly inhibited even in dna+ cells and was nearly completely blocked in dnaB or dnaC(D) mutant. At 37 degrees C, the stage II replication proceeded normally in dna+ bacteria.  相似文献   

15.
A cold-sensitive (CS) mutant of the psychrotroph, Bacillus psychrophilus, was obtained by N-methyl-N'-nitro-N-nitrosoguanidine mutagenization and penicillin counterselection. In the presence of citrate, the wild-type grew well at both 5 and 20 degrees C whereas the CS mutant grew well at 20 degrees C (the permissive temperature) but, at 5 degrees C (the restrictive temperature), grew at a reduced rate for two to three generations followed by a complete plateau in growth. Upon return of the CS mutant to 20 degrees C, after a delay of about 40 h, growth resumed at the appropriate rate. The CS mutant exhibited growth rates similar to parental rates on a wide variety of carbon sources at 5 degrees C, but when Krebs cycle intermediates were used as substrates and in the presence of an equimolar amount of citrate, the typical cold-sensitive growth pattern occurred. Comparison of oxidative phosphorylation in the parent and CS mutant indicated that no phosphorylation occurred at 5 degrees C in the CS mutant during the plateau in growth. Examination of the effect of temperature on ATPase activity showed that at 5 degrees C the specific activity of ATPase isolated from the CS mutant grows at 5 degrees C was 15-fold less than the ATPases isolated from wild-type cells grown at either 5 or 20 degrees C and 10.5-fold lower than ATPase from CS mutant cells grown at 20 degrees C. The large reduction in CS mutant ATPase activity at 5 degrees C appears to be at least partly due to an effect on synthesis since citrate did not inhibit preformed ATPase.  相似文献   

16.
In the filamentous fungus Aspergillus nidulans, the multisubunit motor complex cytoplasmic dynein plays essential roles in nuclear migration and septum positioning. The 8 kDa light chain, LC8, the smallest subunit, is conserved among eukaryotic organisms. Besides being a component in the dynein complex, LC8 also interacts with a wide spectrum of mammalian and viral proteins. To date, the function of this small polypeptide is not well understood. To address this issue, we have created a deletion mutation (DeltanudG) at the nudG locus encoding LC8 in A. nidulans. At 42 degrees C, the DeltanudG mutant forms minute colonies lacking asexual reproduction: this phenotype resembles the phenotype of the dynein heavy chain null mutant. The mutant nuclei largely clustered in the spore body after conidial germination, and the septum was often assembled distally toward the hyphal apex, whereas a control germling has its nuclei distributed along the hypha and the septum formed near the spore body. When the mutant was grown at 23 degrees C, however, its colony resembled a control one, and so did the patterns of nuclear distribution and septum positioning. Elevation of the growth temperature gradually reduced colony size and abolished asexual sporulation. After a period of growth at 23 degrees C that allowed the nuclei to move out of the spore end, a temperature shift to 42 degrees C prevented newly divided nuclei from migrating apart, suggesting that LC8/NUDG was required for both initiating and maintaining dynein motor functions at elevated temperatures. A functional GFP-NUDA fusion was used to test whether LC8/NUDG is required for DHC (dynein heavy chain)/NUDA localization. We found that at 23 degrees C GFP-NUDA localized to the hyphal apex and the septation site in DeltanudG cells as in control cells. Such localizations were absent at 42 degrees C in mutant cells, but not in control cells. We conclude that LC8 plays a role in DHC localization/function, and the requirement for such a role in A. nidulans cells is temperature dependent.  相似文献   

17.
Secretory vesicles that accumulate in the temperature-sensitive sec6-4 strain of yeast have been shown to contain a vanadate-sensitive ATPase, presumably en route to the plasma membrane (Walworth, N. C., and Novick, P. J. (1987) J. Cell Biol. 105, 163-174). We have now established this enzyme to be a fully functional form of the PMA1 [H+]ATPase, identical in its catalytic properties to that found in the plasma membrane. In addition, the secretory vesicles are sealed tightly enough to permit the measurement of ATP-dependent proton pumping with fluorescent probes. We have gone on to exploit the vesicles as an expression system for site-directed mutants of the ATPase. For this purpose, a sec6-4 strain has been constructed in which the chromosomal PMA1 gene is under control of the GAL1 promoter; the mutant pma1 allele to be studied is introduced on a centromeric plasmid under the control of a novel heat shock promoter. In galactose medium at 23 degrees C, the wild-type ATPase is produced and supports normal vegetative growth. When the cells are switched to glucose medium at 37 degrees C, however, the wild-type gene turns off, the mutant gene turns on, and secretory vesicles accumulate. The vesicles contain a substantial amount of newly synthesized, plasmid-encoded ATPase (5-10% of total vesicle protein), but only traces of residual wild-type PMA1 ATPase and no detectable mitochondrial ATPase, vacuolar ATPase, or acid or alkaline phosphatase. To test the expression strategy, we have made use of pma1-105 (Ser368----Phe), a vanadate-resistant mutant previously characterized by standard methods (Perlin, D. S., Harris, S. L., Seto-Young, D., and Haber, J. E. (1989) J. Biol. Chem. 264, 21857-21864). In secretory vesicles, as expected, the plasmid-borne pma1-105 allele gives rise to a mutant enzyme with a reduced rate of ATP hydrolysis and a 100-fold increase in Ki for vanadate. Proton pumping is similarly resistant to vanadate. Thus, the vesicles appear well suited for the production and characterization of mutant forms of the PMA1 [H+]ATPase. They should also aid the study of other yeast membrane proteins that are essential for growth as well as heterologous proteins whose appearance in the plasma membrane may be toxic to the cell.  相似文献   

18.
We report on the properties of a temperature-sensitive mutant produced by transfection of cells with intact DNA and a specific DNA fragment mutagenized with low levels of hydroxylamine. The plating efficiency of the mutant at 39 degrees C relative to that at 33.5 degrees C was 5 X 10(-6). The pattern of polypeptides produced at the nonpermissive temperature was similar to that seen with wild-type virus in infected cells treated with inhibitory concentrations of phosphonoacetic acid in that alpha and beta polypeptides were produced, whereas most gamma polypeptides were either reduced or absent. Consistently, the mutant did not make viral DNA, although temperature sensitivity of the viral DNA polymerase could not be demonstrated. Marker rescue studies with herpes simplex virus type 2 (HSV-2) DNA mapped the mutant in the L component within map positions 0.385 and 0.402 in the prototype (P) arrangement of the HSV-1 genome. Analysis of the recombinants permitted the mapping of the genes specifying infected cell polypeptides 36, 35, 37, 19.5, 11, 8, 2, 43, and 44, but only the infected cell polypeptide 8 of HSV-2 was consistently made by all recombinants containing demonstrable HSV-2 sequences. Marker rescue studies with cloned HSV-1 DNA fragments mapped the temperature-sensitive lesion within less than 10(3) base pairs between 0.383 and 0.388 map units. Translation of the RNA hybridizing to cloned HSV-1 DNA, encompassing the smallest region containing the mutation, revealed polypeptide 8 (128,000 molecular weight), which was previously identified as a beta polypeptide with high affinity for viral DNA, and a polypeptide (25,000 molecular weight) not previously identified in lysates of labeled cells.  相似文献   

19.
The cel mutant of Neurospora, partially blocked in fatty acid synthesis and lacking temperature compensation of its circadian rhythm below 22 degrees C, had a phospholipid fatty acid composition in liquid shaker culture distinctly different from that of a cel+ control strain. During growth, cel+ exhibited a reproducible increase in its linoleic acid level from about 32 to a plateau at 63 mol%, and a corresponding decrease in its linolenic acid level from about 40 to a plateau at 10 mol%. The level of palmitic acid was constant at 19 mol%. In the cel strain, the linoleic acid level was constant at 54 mol% while the palmitic acid level increased from about 12 to about 23 mol%. Supplementation with palmitic or linoleic acids altered the patterns of fatty acid composition of cel, but did not affect the pattern of cel+. Altered fatty acid composition cosegregated with the cel marker. The mitochondrial phospholipids of cel in liquid culture also had abnormal fatty acid composition, as did the whole mycelial phospholipids on solid medium. These results are consistent with the involvement of membrane homeostasis in the temperature compensation of circadian rhythms.  相似文献   

20.
A temperature-conditional mit- mutant of Saccharomyces cerevisiae has been characterized; the mutant strain h45 cannot grow at 36 degrees C on nonfermentable substrates yet appears to be normal at 28 degrees C. The mutation in strain h45 maps genetically to the oli1 region of the mitochondrial DNA (mtDNA) genome, and prevents the synthesis at 36 degrees C of the oli1 gene product, subunit 9 of the mitochondrial ATPase complex. Since the level of oli1 mRNA in mutant h45 is close to normal at 36 degrees C, it is concluded that there is a specific block in translation of this mRNA at the non-permissive temperature. DNA sequence analysis of mtDNA from strain h45 reveals an additional T residue inserted 88 bp upstream of the oli1 coding region, in the A,T-rich sequence that is transcribed into the 5'-untranslated region of the oli1 mRNA. Sequence data on two revertants show that one returns to wild-type parental (J69-1B) mtDNA sequence, whilst the other contains an inserted A residue adjacent to the T inserted in the original h45 mutant. The results are discussed in terms of the stability of folds in RNA upstream of putative ribosome-binding sites in mitochondrial mRNA, and the potential action of nuclear-coded proteins that might be activators of the translation of specific mitochondrial mRNAs in yeast mitochondria.  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