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 31 毫秒
1.
Kinetics of Mitochondrial Complementation   总被引:1,自引:0,他引:1  
Mitochondrial complementation (enhancement of oxidative activity of mitochondrial occured when mitochondria of some inbrveds of eron were mixed in vitro. It was found that mixtures composed of various proportions of mitochondria of inbreds exhibited complementation. Extracts of mitochondria of one inbred did not promote complementation when they were added to infact mitochondria of inbreds exhihited complententation. Extracts of mitochondria of on einbred did not promote complementation when they were added to infact mitochondria of the second inbred. Serial dillutions of mixtures resulted in a rapid reduction of mitochondrial oxidation (QO2 N) suggesting that intimale association between mitochondria was minimized by the dilution and that this decreased complementation. Similar dilutions of mitochondria from inbreds did not decrease their specific activity. Kinetics of mitochondrial oxidation were measured polarographically and it was found that complementation was measurable immediately (1 min) following the mixing. This suggested that physical contact between mitochondria was necessary for complementation.  相似文献   

2.
Four temperature-sensitive HPRT clones were used for hybridological analysis, which led to increase in complementation rate about 5 times. The probability of complementation, in respect of the HPRT locus proved to be rather high: 14 of 45 hybridization-tested mutants had complementation ability (including 3 ts mutants). Analysis of the complementation rate among mutants revealed clear-cut dependence on the selection conditions: clones grown in a medium with 8-azaguanine showed most frequent complementation. The use of mutants with a new phenotype in hybridization analysis revealed four additional complementation groups, three of which are made of temperature-sensitive clones. Biochemical analysis revealed the presence of hybrid forms of the HPRT enzyme in all hybrids tested. This confirms the intragenic character of complementation. At present, the functional map of the HPRT locus is represented by 9 groups, including a group of mutants with no complementation ability.  相似文献   

3.
Temperature-sensitive mutants of Sindbis virus fail to grow at a temperature that permits growth of the wild type, but when certain pairs of these mutants, mixed together, infect cells at that temperature, viral growth (i.e., complementation) occurs. The yield from this complementation, however, is of the same order of magnitude as the infectivity in the inoculum. Since in animal virus infections the protein components of the virion probably enter the cell with the viral nucleic acid, it was necessary to demonstrate that the observed complementation required synthesis of new viral protein and nucleic acid rather than some sort of rearrangement of the structural components of the inoculum. To demonstrate that complementation does require new biosynthesis, three biochemical events of normal virus growth have been observed during complementation and correlated with the efficiency of viral growth seen in complementation. These events include: (i) entrance of parental viral ribonucleic acid (RNA) into a double-stranded form; (ii) subsequent synthesis of viral RNA; and (iii) synthesis and subsequent incorporation of viral protein(s) into cell membranes where they were detected by hemadsorption. Although the infecting single-stranded RNA genome of the wild type was converted to a ribonuclease-resistant form, the genome of a mutant (ts-11) incapable of RNA synthesis at a nonpermissive temperature was not so converted. However, during complementation with another mutant also defective in viral RNA synthesis, some of the RNA of mutant ts-11 was converted to a ribonuclease-resistant form, and total synthesis of virus-specific RNA was markedly enhanced. The virus-specific alteration of the cell surface, detected by hemadsorption, was also extensively increased during complementation. These observations support the view that complementation between temperature-sensitive mutants and replication of wild-type virus are similar processes.  相似文献   

4.
Korch CT 《Genetics》1973,74(2):307-329
Sixty-two alleles of the histidine-1 (his1) gene were tested for complementation. The 44 complementing mutants fell into 31 complementation groups which were used to construct a complex complementation map with 18 complementation units. Cluster analysis of the complementation map by either visual inspection or the computer method of Gillie and Peto (1969) shows two very definite clusters.The molecular weight estimate of the his1 enzyme, phosphoribosyl adenosine triphosphate: pyrophosphate phosphoribosyltransferase, is 1.8 . 10(5) by sucrose density gradient analysis and 2.4 . 10(5) by Sephadex gel chromatography. Correlating the length of the his1 gene to the molecular weight of the enzyme indicates that this enzyme is composed of 6 subunits, as is the analogous enzyme in Salmonella typhimurium.A model of the subunit and tertiary and quaternary structure of the enzyme has been developed from consideration of the genetic and complementation data, the distribution of the various mutant types within the gene, and the biochemical properties of the enzyme encoded by the his1 gene.  相似文献   

5.
We have isolated mutants sensitive to methyl methanesulfonate (MMS) in Saccharomyces cerevisiae. Alleles of rad1, rad4, rad52, rad55 and rad57 were found amoung these mms mutants. Twenty-nine of the mms mutants which complement the existing radiation-sensitive (rad and rev) mutants belong to 22 new complementation groups. Mutants from five complementation groups are sensitive only to MMS. Mutants of 11 complementation groups are sensitive to UV or X rays in addition to MMS, mutants of six complementation groups are sensitive to all three agents. The cross-sensitivities of these mms mutants to UV and X rays are discussed in terms of their possible involvement in DNA repair. Sporulation is reduced or absent in homozygous diploids of mms mutants from nine complementation groups.  相似文献   

6.
Sixteen temperature-sensitive mutants of Sendai virus were isolated from mutagenized stocks (10 mutants, designated numerically) and persistently infected cultures (6 mutants, designated alphabetically). Based on complementation tests, virion-associated activities, thermal inactivation, and viral RNA and hemadsorbing antigen synthesis as well as virion production in chick lung embryo cells at nonpermissive temperature, these mutants were divided into seven groups as follows. i) HANA group mutants (ts-5, -9, -10, -201), defective in hemagglutinin-neuraminidase protein, complementation group I. ii) F group mutants (ts-18, -108), defective in hemolytic and cell-fusing activity, complementation group II. iii) Ts-43, defective in RNA polymerase activity, complementation group III. iv) Ts-23, defective in RNA polymerase activity, interfered with the other mutants in complementation tests. v) Ts-25, defective in the incorporation of hemagglutinin-neuraminidase protein into the virion at the stage of virus assembly. vi) Ts-110, belongs to F group mutants on one hand, but is considered to carry another undetermined defect. vii) C group (carrier culture-borne group) mutants (ts-a, -b, -c, -d, -e, -f), defective lesion not yet determined and belong to neither complementation group I nor II. Assignment of mutants in groups iv), v), vi), and vii) to complementation groups could not be achieved.  相似文献   

7.
Amber mutants of bacteriophage BF23 were classified into two functional groups, types I and II, by the yields of the infecting-mutant genotypes in plate complementation tests. Type I mutants produced their genotypes at levels more than 20% of the total progeny phages, and type II mutants did so at levels of less than 5%. Comparison of the results of plate complementation tests with those of extract complementation tests revealed that all the type I mutants were defective in the tail formation, while most type II mutants were defective in the formation of either mature heads (type IIa) or both mature heads and tails (type IIb). Since in extract complementation tests the activated phages are always of genotypes corresponding to mutations defective in only the tail formation, the plate complementation test is comparable with the extract complementation test when judged on the basis of the yield of the mutant genotypes. Of 29 complementation groups, 8 type I, 14 type IIa, and 5 type IIb mutants were identified. Previously, amber mutations of BF23 were mapped on four genetic segments. These segments were ordered in one linkage map by crosses between deletion and amber mutants.  相似文献   

8.
We have conducted a genetic analysis of the region flanking the 68C glue gene cluster in Drosophila melanogaster by isolating lethal and semilethal mutations uncovered by deficiencies which span this region. Three different mutagens were used: ethyl methanesulfonate (EMS), ethyl nitrosourea (ENU) and diepoxybutane (DEB). In the region from 68A3 to 68C11, 64 lethal, semilethal, and visible mutations were recovered. These include alleles of 13 new lethal complementation groups, as well as new alleles of rotated, low xanthine dehydrogenase, lethal(3)517 and lethal(3)B76. Six new visible mutations from within this region were recovered on the basis of their reduced viability; all proved to be semiviable alleles of lethal complementation groups. No significant differences were observed in the distributions of lethals recovered using the three different mutagens. Each lethal was mapped on the basis of complementation with overlapping deficiencies; mutations that mapped within the same interval were tested for complementation, and the relative order of the lethal groups within each interval was determined by recombination. The cytological distribution of genes within the 68A3-68C11 region is not uniform: the region from 68A2,3 to 68B1,3 (seven to ten polytene chromosome bands) contains at least 13 lethal complementation groups and the mutation low xanthine dehydrogenase; the adjoining region from 68B1,3 to 68C5,6 (six to nine bands) includes the 68C glue gene cluster, but no known lethal or visible complementation groups; and the interval from 68C5,6 to 68C10,11 (three to five bands) contains at least three lethal complementation groups and the visible mutation rotated. The developmental stage at which lethality is observed was determined for a representative allele from each lethal complementation group.  相似文献   

9.
Complementation studies, using fused cell lines from patients with peroxisomal disorders, have shown correction of defective plasmalogen synthesis and phytanic acid oxidation as well as an increase in the number of peroxisomes. At least six complementation groups have been reported. We demonstrate here that complementing cell lines also acquire the ability to oxidize very long chain fatty acids (VLCFA), and that complementation groups defined with this technique are identical to those reported previously when plasmalogen synthesis was used as the criterion for complementation. This VLCFA complementation technique is of particular value in the study of patients in whom defective VLCFA is the only or major enzymatic defect, and we show complementation between cell lines from two patients each with an isolated defect in one of the peroxisomal fatty acid beta-oxidation enzymes.  相似文献   

10.
Free-living birds must satisfy fluctuating nutrient requirements in diverse and varying environments. Ingesting nutritionally complementary foods may be the most effective means by which wild birds match nutrient ingestion and nutrient needs. Dietary complementation may occur fortuitously when foods chosen in response to non-nutritive factors (e.g. competition, predation risk, food colour), or on the basis of energy density, also fulfill specific nutrient needs (passive dietary complementation). In some environments, especially during productive phases (e.g. reproduction), free-living birds may rely on nutrient appetites to ensure their choice of foods satisfies their nutrient needs (active dietary complementation). Meeting nutrient needs through dietary complementation can be facilitated, complicated, or impeded by any of several environmental or organism determinants of food choice. Nutrient appetites, exogenous food stores, and endogenous nutrient stores are three organismal determinants that are probably the most important in facilitating dietary complementation.  相似文献   

11.
Twelve temperature-sensitive (ts) mutants of herpes simplex virus type 1 (HSV-1), representing seven complementation groups, were isolated subsequent to 5-bromodeoxyuridine mutagenesis. These mutants were identified by their inability to replicate in a line of monkey (CV-1) cells at 39 C. Seven of these mutants, representing six complementation groups, induced thymidine kinase (tk) and transformed Ltk- cells, a line of mouse L cells lacking tk, to a tk+ phenotype at both the permissive (34 C) and nonpermissive (39 C) temperatures. Thus, the defective cistrons in these six complementation groups, although necessary for lysis, have no essential function in this transformation system. Transformation by these 12 mutants was dependent on prior UV irradiation. Infection of cells with unirradiated virus under conditions which did not permit virus replication was not sufficient to allow cell transformation. Five mutants, representing two complementation groups, were tk- and were incapable of causing the tk--to-tk+ transformation at either 34 C of 39 C. The tk defects in these mutants are probably unrelated to the ts defects, since one of these complementation groups contains a tk+ member. Therefore, transformation of Ltk- cells to a tk+ phenotype by HSV-1 requires an active viral tk gene. One complementation group was represented by a single tk- member. The role of this cistron in transformation remains undetermined since the primary block to transformation is presumed to be the tk- phenotype. Mutants representing the seven complementation groups were unable to replicate at 39 C in two lines of HSV-1-transformed cells, indicating that the activities of resident wild-type copies of the defective cistrons, if present, could not be detected by complementation.  相似文献   

12.
Tzeng WP  Frey TK 《Journal of virology》2003,77(17):9502-9510
Rubella virus (RUB) replicons with an in-frame deletion of 507 nucleotides between two NotI sites in the P150 nonstructural protein (DeltaNotI) do not replicate (as detected by expression of a reporter gene encoded by the replicon) but can be amplified by wild-type helper virus (Tzeng et al., Virology 289:63-73, 2001). Surprisingly, virus with DeltaNotI was viable, and it was hypothesized that this was due to complementation of the NotI deletion by one of the virion structural protein genes. Introduction of the capsid (C) protein gene into DeltaNotI-containing replicons as an in-frame fusion with a reporter gene or cotransfection with both DeltaNotI replicons and RUB replicon or plasmid constructs containing the C gene resulted in replication of the DeltaNotI replicon, confirming the hypothesis that the C gene was the structural protein gene responsible for complementation and demonstrating that complementation could occur either in cis or in trans. Approximately the 5' one-third of the C gene was necessary for complementation. Mutations that prevented translation of the C protein while minimally disturbing the C gene sequence abrogated complementation, while synonymous codon mutations that changed the C gene sequence without affecting the amino acid sequence at the 5' end of the C gene had no effect on complementation, indicating that the C protein, not the C gene RNA, was the moiety responsible for complementation. Complementation occurred at a basic step in the virus replication cycle, because DeltaNotI replicons failed to accumulate detectable virus-specific RNA.  相似文献   

13.
Nine lethal complementation groups flanking the Drosophila Dopa decarboxylase (Ddc) gene, have been localized within 100 kb of cloned chromosomal DNA. Six of these complementation groups are within 23 kb of DNA, and all ten complementation groups, including Ddc, lie within 78-82 kb of DNA. The potential significance of this unusually high gene density is discussed.  相似文献   

14.
Rhizobium meliloti produces an acidic, Calcofluor-binding exopolysaccharide which plays a role in nodulation of alfalfa plants by this bacterium. We constructed and mapped 102 transposon insertions in a 48-kilobase (kb) region previously shown to contain several exo genes. Mutations affecting production of the Calcofluor-binding exopolysaccharide were clustered in a 22-kb region and fell into 12 complementation groups. Strains carrying mutations in seven of the complementation groups (exoA, exoB, exoF, exoL, exoM, exoP, and exoQ) produced no Calcofluor-binding exopolysaccharide and induced non-nitrogen-fixing nodules on alfalfa. Mutants in an eighth complementation group, exoH (Leigh et al., Cell 51:579-587, 1987), produce an altered exopolysaccharide and also induce the formation of non-nitrogen-fixing nodules. Mutants in the remaining four complementation groups produced less Calcofluor-binding material than the wild type. Mutants carrying mutations in two of these complementation groups (exoK and exoN) formed apparently normal, nitrogen-fixing nodules, while mutants in the other two groups (exoG and exoJ) formed normal nodules less efficiently than the wild type.  相似文献   

15.
Fanconi anemia (FA) is an autosomal recessive chromosomal instability syndrome with at least seven different complementation groups. Four FA genes (FANCA, FANCC, FANCF, and FANCG) have been identified, and two other FA genes (FANCD and FANCE) have been mapped. Here we report the identification, by complementation cloning, of the gene mutated in FA complementation group E (FANCE). FANCE has 10 exons and encodes a novel 536-amino acid protein with two potential nuclear localization signals.  相似文献   

16.
We studied genetic complementation of propionyl CoA carboxylase (PCC) deficiency in cultures of polyethylene glycol (PEG)-induced heterokaryons, using mutant fibroblast lines assigned to five mutant classes, designated bio, pcc A, pcc B, pcc C, and pcc BC. By measuring PCC activity directly in extracts of fused cells or indirectly in intact cells by [1-14C]propionate utilization, we confirmed the nonlinear nature of the PCC deficiency complementation map described by Gravel et al. [1]. When we studied the kinetics of complementation, we detected three distinct patterns using the [1-14C]propionate utilization assay. When either pcc A or pcc C lines were fused to bio cells, 14C-fixation increased to half of the maximally restored values within 4 hrs. In pcc A x pcc C crosses or in pcc A x pcc B crosses, however, complementation was much slower. In fusions between pcc B and pcc C cells, a third pattern was elicited; complementation was incomplete, maximum restoration of PCC activity begin less than 20% of that observed in other complementing crosses. From these data and previous biochemical evidence, we suggest (1) that the bio and pcc mutations affect different genes; (2) that complementation between pcc A and either pcc B, pcc C, or pcc BC lines is intergenic and involves subunit exchange and synthesis of new PCC molecules; and (3) that complementation between pcc B and pcc C mutants is interallelic.  相似文献   

17.
V-H4, a mitomycin C (MMC)-sensitive Chinese hamster cell mutant, is phenotypically very similar to Fanconi anemia (FA) cells. Genetic complementation analysis shows that V-H4 belongs to the same complementation group as FA group A cells. Proliferating hybrid cell lines obtained after fusion of V-H4 with normal or FA group B cells show an increased resistance to MMC. Absence of complementation was noted in V-H4 x FA group A hybrid cell lines. This was shown not to be due to the absence of a specific human chromosome. The V-H4 mutant represents the first rodent mutant that is genotypically similar to FA complementation group A cells.  相似文献   

18.
The RAD10 gene of Saccharomyces cerevisiae is required for nucleotide excision repair of DNA. Expression of RAD10 mRNA and Rad10 protein was demonstrated in Chinese hamster ovary (CHO) cells containing amplified copies of the gene, and RAD10 mRNA was also detected in stable transfectants without gene amplification. Following transfection with the RAD10 gene, three independently isolated excision repair-defective CHO cell lines from the same genetic complementation group (complementation group 2) showed partial complementation of sensitivity to killing by UV radiation and to the DNA cross-linking agent mitomycin C. These results were not observed when RAD10 was introduced into excision repair-defective CHO cell lines from other genetic complementation groups, nor when the yeast RAD3 gene was expressed in cells from genetic complementation group 2. Enhanced UV resistance in cells carrying the RAD10 gene was accompanied by partial reactivation of the plasmid-borne chloramphenicol acetyltransferase (cat) gene following its inactivation by UV radiation. The phenotype of CHO cells from genetic complementation group 2 is also specifically complemented by the human ERCC1 gene, and the ERCC1 and RAD10 genes have similar amino acid sequences. The present experiments therefore indicate that the structural homology between the yeast Rad10 and human Ercc1 polypeptides is reflected at a functional level, and suggest that nucleotide excision repair proteins are conserved in eukaryotes.  相似文献   

19.
The multisystem autosomal recessive disease ataxia-telangiectasia (A-T) is determined by several genes, as evidenced by the existence of four complementation groups in this disorder. Using linkage analysis, the ATA (A-T complementation group A) gene was previously localized to chromosome 11, region q22-q23. Analysis of the segregation of RFLP markers from this region in a Jewish-Moroccan family assigned to group C indicates that the ATC (A-T complementation group C) gene localizes to chromosome 11q22-q23 as well.  相似文献   

20.
BACKGROUND: Gene complementation strategies are important in validating the roles of genes in specific phenotypes. Complementation systems in Helicobacter pylori include shuttle vectors, which transform H. pylori at relatively low frequencies, and chromosomally based approaches. Chromosomal complementation strategies are susceptible to polar effects and disruption of other H. pylori genes, leading to unwanted pleiotropic effects. MATERIALS AND METHODS: A new complementation strategy was developed for H. pylori by utilizing a suicide plasmid vector that contains fragments of an H. pylori intergenic region (hp0203-hp0204), a chloramphenicol acetyltransferase cassette (cat), and a multiple-cloning site. Genes of interest could be cloned into the intergenic plasmid and the genes integrated into H. pylori by homologous recombination into the intergenic chromosomal region without disrupting any annotated H. pylori gene. The complementation system was validated using the gene encoding arginase (rocF). RESULTS: A rocF mutant unable to hydrolyze or consume l-arginine regained these functions by complementation with the wild-type rocF gene. Complemented strains also had restored arginase protein as determined by Western blot analysis. The complementation system could be successfully applied to multiple H. pylori strains. The intergenic region varied in length and sequence across 17 H. pylori strains, but the flanking-3' ends of the hp0203 and hp0204 coding regions were highly conserved. Inserting a cat cassette and wild-type rocF into the intergenic region did not alter the ability of strain SS1 to colonize mice. CONCLUSIONS: This complementation strategy should greatly facilitate genetic experiments in H. pylori.  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