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.
Posterior polymorphous corneal dystrophy (PPCD) is a rare autosomal dominant genetically heterogeneous disorder. Nineteen Czech PPCD pedigrees with 113 affected family members were identified, and 17 of these kindreds were genotyped for markers on chromosome 20p12.1- 20q12. Comparison of haplotypes in 81 affected members, 20 unaffected first degree relatives and 13 spouses, as well as 55 unrelated controls, supported the hypothesis of a shared ancestor in 12 families originating from one geographic location. In 38 affected individuals from nine of these pedigrees, a common haplotype was observed between D20S48 and D20S107 spanning approximately 23 Mb, demonstrating segregation of disease with the PPCD1 locus. This haplotype was not detected in 110 ethnically matched control chromosomes. Within the common founder haplotype, a core mini-haplotype was detected for D20S605, D20S182 and M189K2 in all 67 affected members from families 1–12, however alleles representing the core mini-haplotype were also detected in population matched controls. The most likely location of the responsible gene within the disease interval, and estimated mutational age, were inferred by linkage disequilibrium mapping (DMLE+2.3). The appearance of a disease-causing mutation was dated between 64–133 generations. The inferred ancestral locus carrying a PPCD1 disease-causing variant within the disease interval spans 60 Kb on 20p11.23, which contains a single known protein coding gene, ZNF133. However, direct sequence analysis of coding and untranslated exons did not reveal a potential pathogenic mutation. Microdeletion or duplication was also excluded by comparative genomic hybridization using a dense chromosome 20 specific array. Geographical origin, haplotype and statistical analysis suggest that in 14 unrelated families an as yet undiscovered mutation on 20p11.23 was inherited from a common ancestor. Prevalence of PPCD in the Czech Republic appears to be the highest worldwide and our data suggests that at least one other novel locus for PPCD also exists.  相似文献   

2.

Background

Several founder mutations leading to increased risk of cancer among Ashkenazi Jewish individuals have been identified, and some estimates of the age of the mutations have been published. A variety of different methods have been used previously to estimate the age of the mutations. Here three datasets containing genotype information near known founder mutations are reanalyzed in order to compare three approaches for estimating the age of a mutation. The methods are: (a) the single marker method used by Risch et al., (1995); (b) the intra-allelic coalescent model known as DMLE, and (c) the Goldgar method proposed in Neuhausen et al. (1996), and modified slightly by our group. The three mutations analyzed were MSH2*1906 G->C, APC*I1307K, and BRCA2*6174delT.

Results

All methods depend on accurate estimates of inter-marker recombination rates. The modified Goldgar method allows for marker mutation as well as recombination, but requires prior estimates of the possible haplotypes carrying the mutation for each individual. It does not incorporate population growth rates. The DMLE method simultaneously estimates the haplotypes with the mutation age, and builds in the population growth rate. The single marker estimates, however, are more sensitive to the recombination rates and are unstable. Mutation age estimates based on DMLE are 16.8 generations for MSH2 (95% credible interval (13, 23)), 106 generations for I1037K (86-129), and 90 generations for 6174delT (71-114).

Conclusions

For recent founder mutations where marker mutations are unlikely to have occurred, both DMLE and the Goldgar method can give good results. Caution is necessary for older mutations, especially if the effective population size may have remained small for a long period of time.
  相似文献   

3.
In Escherichia coli the codons CGG, AGG, UGG or GGG (NGG codons) but not GGN or GNG (where N is non-G) are associated with low expression of a reporter gene, if located at positions +2 to +5. Induction of a lacZ reporter gene with any one of the NGG codons at position +2 to +5 does not influence growth of a normal strain, but growth of a strain with a defective peptidyl-tRNA hydrolase (Pth) enzyme is inhibited. The same codons, if placed at position +7, did not give this effect. Other codons, such as CGU and AGA, at location +2 to +5, did not give any growth inhibition of either the wild-type or the mutant strain. The inhibitory effect on the pth mutant strain by NGG codons at location +5 was suppressed by overexpression of the Pth enzyme from a plasmid. However, the overexpression of cognate tRNAs for AGG or GGG did not rescue from the growth inhibition associated with these codons early in the induced model gene. The data suggest that the NGG codons trigger peptidyl-tRNA drop-off if located at early coding positions in mRNA, thereby strongly reducing gene expression. This does not happen if these codons are located further down in the mRNA at position +7, or later.  相似文献   

4.
The envB1 mutation mediating a distorted cell morphology of Escherichia coliK-12 was cotransducible with strA, aroE, aspB, and argG. The mapping data is consistent with a gene location for envB around 62.5 min. In partial diploids envB1 was recessive to its wild-type allele. The original envB mutant contained a second mutation in a locus denoted sloB close to strA. The following gene order is suggested: sloB-strA-aroE-envB-aspB-argG. The sloB1 mutation caused a marked reduction in the growth rate of both envB and envB+ strains. Moreover, this mutation in the presence of envB1 appears to increase the ratio between deoxyribonucleic acid and protein in cells growing in rich medium. The phenotypic properties of envB1, sloB+ and envB+ transductants were characterized. Cells with envB1, sloB+ genotype were hypersensitive to several penicillins including the beta-lactam compound, amidino penicillin. Penicillin hypersensitivity could not be explained by increased outer membrane penetrability. The original envB mutant (envB1,SLOB1), as well as envB1, sloB1 or envB+, SLOB1 transductants were resistant to amidino penicillin. Resistance was explained by the slow growth rate mediated by the sloB1 mutation. The similarity between envB cells and wild-type cells treated with sublethal concentrations of amidino penicillin was emphasized.  相似文献   

5.
A new method is presented for fine-scale linkage disequilibrium (LD) mapping of a disease mutation; it uses multiple linked single-nucleotide polymorphisms, restriction-fragment-length polymorphisms, or microsatellite markers and incorporates information from an annotated human genome sequence (HGS) and from a human mutation database. The method takes account of population demographic effects, using Markov chain Monte Carlo methods to integrate over the unknown gene genealogy and gene coalescence times. Information about the relative frequency of disease mutations in exons, introns, and other regions, from mutational databases, as well as assumptions about the completeness of the gene annotation, are used with an annotated HGS, to generate a prior probability that a mutation lies at any particular position in a specified region of the genome. This information is updated with information about mutation location, from LD at a set of linked markers in the region, to generate the posterior probability density of the mutation location. The performance of the method is evaluated by simulation and by analysis of a data set for diastrophic dysplasia (DTD) in Finland. The DTD disease gene has been positionally cloned, so the actual location of the mutation is known and can be compared with the position predicted by our method. For the DTD data, the addition of information from an HGS results in disease-gene localization at a resolution that is much higher than that which would be possible by LD mapping alone. In this case, the gene would be found by sequencing a region < or =7 kb in size.  相似文献   

6.
A hybrid hsdS gene, encoding the HsdSts + d polypeptide, was constructed by joining the proximal region of the wild-type (wt) hsdS sequence with the distal region of the hsdSts + d sequence, at the hsdS BglII site. The hybrid hsdS-Sts + d gene exerts a trans-dominant effect on restriction and modification, which points to the location of the temperature-sensitive (ts) trans-dominant (+ d) mutation in the gene hsdSts + d distal region. Sequencing of the region downstream from the HindIII target in the Escherichia coli K-12 hsdSts + d mutant was carried out. It is identical to the wt hsdS sequence (GenBank/EMBL accession number ECHSDK LV00288), except for a single base-pair transition C1245----T. The results obtained support the idea that the trans-dominant effect of the ts mutation described earlier is related to the single base-pair transition in the nonhomologous region of the hsdSts + d sequence.  相似文献   

7.
When the Lac- strain of Escherichia coli, FC40, is incubated with lactose as its sole carbon and energy source, Lac+ revertants arise at a constant rate, a phenomenon known as adaptive mutation. Two alternative models for adaptive mutation have been proposed: (i) recombination-dependent mutation, which specifies that recombination occurring in nongrowing cells stimulates error-prone DNA synthesis, and (ii) amplification-dependent mutation, which specifies that amplification of the lac region and growth of the amplifying cells creates enough DNA replication to produce mutations at the normal rate. Here, we examined several of the predictions of the amplification-dependent mutation model and found that they are not fulfilled. First, inhibition of adaptive mutation by a gene that is toxic when overexpressed does not depend on the proximity of the gene to lac. Second, mutation at a second locus during selection for Lac+ revertants is also independent of the proximity of the locus to lac. Third, mutation at a second locus on the episome occurs even when the lac allele under selection is on the chromosome. Our results support the hypothesis that most Lac+ mutants that appear during lactose selection are true revertants that arise in a single step from Lac- cells, not from a population of growing or amplifying precursor cells.  相似文献   

8.
Leesch VW  Bujons J  Mauk AG  Hoffman BM 《Biochemistry》2000,39(33):10132-10139
Cytochrome c peroxidase (CcP) can bind as many as two cytochrome c (Cc) molecules in an electrostatic complex. The location of the two binding domains on CcP has been probed by photoinduced interprotein electron transfer (ET) between zinc-substituted horse cytochrome c (ZnCc) and CcP with surface charge-reversal mutations and by isothermal titration calorimetry (ITC). These results, which are the first experimental evidence for the location of domain 2, indicate that the weak-binding domain includes residues 146-150 on CcP. CcP(E290K) has a charge-reversal mutation in the tight-binding domain, which should weaken binding, and it weakens the 1:1 complex; K(1) decreases 20-fold at 18 mM ionic strength. We have employed two mutations to probe the proposed location for the weak-binding domain on the CcP surface: (i) D148K, a "detrimental" mutation with a net (+2) change in the charge of CcP, and (ii) K149E, a "beneficial" mutation with a net (-2) change in the charge. The interactions between FeCc and CcP(WT and K149E) also have been studied with ITC. The CcP(D148K) mutation causes no substantial change in the 2:1 binding but an increase in the reactivity of the 2:1 complex. The latter can be interpreted as a long-range influence on the heme environment or, more likely, the enhancement of a minority subset of binding conformations with favorable pathways for ET. CcP(K149E) has a charge-reversal mutation in the weak-binding domain that produces a substantial increase in the 2:1 binding constant as measured by both quenching and ITC. For the 1:1 complex of CcP(WT), DeltaG(1) = -8.2 kcal/mol (K(1) = 1.3 x 10(6) M(-)(1)), DeltaH(1) = +2.7 kcal/mol, and DeltaS(1) = +37 cal/K.mol at 293 K; for the second binding stage, K(2) < 5 x 10(3) M(-)(1), but accurate thermodynamic parameters were not obtained. For the 1:1 complex of CcP(K149E), DeltaG(1) = -8.5 kcal/mol (K(1) = 2 x 10(6) M(-)(1)), DeltaH(1) = +2. 0 kcal/mol, and DeltaS(1) = +36 cal/K.mol; for the second stage, DeltaG(2) = -5.5 kcal/mol (K(1) = 1.3 x 10(4) M(-)(1)), DeltaH(2) = +2.9 kcal/mol, and DeltaS(2) = +29 cal/K.mol.  相似文献   

9.
The Phox2b gene is necessary for autonomic nervous-system development. Phox2b-/- mice die in utero with absent autonomic nervous system circuits, since autonomic nervous system neurons either fail to form or degenerate. We first identified the Phox2b human ortholog, PHOX2B, as the gene underlying congenital central hypoventilation syndrome (CCHS, or Ondine curse), with an autosomal dominant mode of inheritance and de novo mutation at the first generation. We have subsequently shown that heterozygous mutations of PHOX2B may account for several combined or isolated disorders of autonomic nervous-system development--namely, tumors of the sympathetic nervous system (TSNS), such as neuroblastoma and late-onset central hypoventilation syndrome. Here, we report the clinical and molecular assessments of a cohort of 188 probands with CCHS, either isolated or associated with Hirschsprung disease and/or TSNS. The mutation-detection rate was 92.6% (174/188) in our series, and the most prevalent mutation was an in-frame duplication leading to an expansion of +5 to +13 alanines in the 20-alanine stretch at the carboxy terminal of the protein. Such findings suggest PHOX2B mutation screening as a simple and reliable tool for the diagnosis of CCHS, independent of the clinically variable phenotype. In addition, somatic mosaicism was detected in 4.5% of parents. Most interestingly, analysis of genotype-phenotype interactions strongly supports the contention that patients with CCHS who develop malignant TSNS will harbor either a missense or a frameshift heterozygous mutation of the PHOX2B gene. These data further highlight the link between congenital malformations and tumor predisposition when a master gene in development is mutated.  相似文献   

10.
Mutations in recA, such as recA801(Srf) (suppressor of RecF) or recA441(Tif) (temperature-induced filamentation) partially suppress the deficiency in postreplication repair of UV damage conferred by recF mutations. We observed that spontaneous recA(Srf) mutants accumulated in cultures of recB recC sbcB sulA::Mu dX(Ap lac) lexA51 recF cells because they grew faster than the parental strain. We show that in a uvrA recB+ recC+ genetic background there are two prerequisites for the suppression by recA(Srf) of the UV-sensitive phenotype of recF mutants. (i) The recA(Srf) protein must be provided in increased amounts either by SOS derepression or by a recA operator-constitutive mutation in a lexA(Ind) (no induction of SOS functions) genetic background. (ii) The gene recJ, which has been shown previously to be involved in the recF pathway of recombination and repair, must be functional. The level of expression of recJ in a lexA(Ind) strain suffices for full suppression. Suppression by recA441 at 30 degrees C also depends on recJ+. The hampered induction by UV of the SOS gene uvrA seen in a recF mutant was improved by a recA(Srf) mutation. This improvement did not require recJ+. We suggest that recA(Srf) and recA(Tif) mutant proteins can operate in postreplication repair independent of recF by using the recJ+ function.  相似文献   

11.
gltBDF operon of Escherichia coli.   总被引:14,自引:10,他引:4       下载免费PDF全文
A 2.0-kilobase DNA fragment carrying antibiotic resistance markers was inserted into the gltB gene of Escherichia coli previously cloned in a multicopy plasmid. Replacement of the chromosomal gltB+ gene by the gltB225::omega mutation led to cells unable to synthesize glutamate synthase, utilize growth rate-limiting nitrogen sources, or derepress their glutamine synthetase. The existence of a gltBDF operon encoding the large (gltB) and small (gltD) subunits of glutamate synthase and a regulatory peptide (gltF) at 69 min of the E. coli linkage map was deduced from complementation analysis. A plasmid carrying the entire gltB+D+F+ operon complemented cells for all three of the mutant phenotypes associated with the polar gltB225::omega mutation in the chromosome. By contrast, plasmids carrying gltB+ only complemented cells for glutamate synthase activity. A major tricistronic mRNA molecule was detected from Northern (RNA blot) DNA-RNA hybridization experiments with DNA probes containing single genes of the operon. A 30,200-dalton polypeptide was identified as the gltF product, the lack of which was responsible for the inability of cells to use nitrogen-limiting sources associated with gltB225::omega.  相似文献   

12.
M. Lieb 《Genetics》1991,128(1):23-27
In many strains of Escherichia coli, the product of gene dcm methylates the internal cytosines in the sequence 5'CC(A or T)GG. Spontaneous deamination of 5-methylcytosine produces thymine which, if not corrected, can result in a transition mutation. 5-Methylcytosines in the lacI gene are hotspots for spontaneous C to T mutations. dcm is linked to vsr, a gene required for very short patch (VSP) repair. VSP repair corrects T.G mispairs in the following contexts:CTAAGGGGTCC, CTTGGGGACC, TAGGGTCC and CTAGGGTC. I have investigated the relationships between cytosine methylation, mutation, and VSP repair. Spontaneous mutations in the repressor (cI) gene of lambda prophage were isolated in wild-type and mutant lysogens. A hotspot for spontaneous mutation that corresponds with a 5-methylcytosine was observed in wild-type lysogens but was not present in bacteria lacking both methylase and VSP repair activity. Introduction of a plasmid containing dcm+ and vsr+ restored the mutation hotspot. If the added plasmid carried only dcm+, the frequency of spontaneous mutations at the 5-methylcytosine was over 10-fold higher than in Dcm+Vsr+ lysogens. The addition of vsr on a plasmid to a wild-type lysogen resulted in a 4-fold reduction in mutation at the hotspot. These findings support the previously untested hypothesis that VSP repair prevents mutations resulting from deamination of 5-methylcytosine.  相似文献   

13.
The pcsA gene is identical to dinD in Escherichia coli.   总被引:3,自引:1,他引:2       下载免费PDF全文
The pcsA68 mutant of Escherichia coli is a cold-sensitive mutant which forms long filaments with a large nucleoid in the central region at 20 degrees C. We here show that (i) the coding region for the pcsA gene is identical with orfY located upstream of pyrE and can be deleted without loss of viability; (ii) pcsA is also identical to dinD, a DNA damage-inducible gene, whose expression is regulated by the LexA-RecA system; (iii) the cold-sensitive phenotype of the pcsA68 mutation is suppressed by delta recA or lexA1 (Ind-) mutation, but not by sulA inactivation; (iv) overproduction of PcsA68 leads to inhibition of cell growth in recA+ and delta recA strains at 20 and 37 degrees C, but PcsA+ does not show such an effect at any temperature; (v) SOS response is induced in the pcsA68 mutant cells at 20 degrees C. We discuss the possible function of the pcsA gene, comparing it with the sulA or the dif-xerCD function. We also describe a new method for gene disruption with positive and negative selection.  相似文献   

14.
tolA mutants of Escherichia coli K-12 release periplasmic proteins into the extracellular medium; they are sensitive to growth inhibitors such as cholic acid and tolerant to group A colicins and filamentous bacteriophage. Suppressor mutants of the tolA-876 allele were isolated by selecting for cholic acid resistant clones that did not release periplasmic ribonuclease I. One class of tolA suppressor strains carried mutations in the staA gene (for suppressor of tolA) located a 41 min. tolA-876 staA strains partially recovered a wild-type phenotype: they exported alkaline phosphatase and beta-lactamase into the periplasm and only released very low amounts of periplasmic proteins; moreover, they were sensitive to E1 and A colicins and more resistant than tolA-876 staA+ strains to various growth inhibitors. Furthermore, tolA-876 staA-2 and tolA+staA-2 mutants were 10- to 2700-times more resistant than staA+ strains to bacteriophages TuIa, TuIb and T4, and TuII whose receptors are major outer membrane proteins OmpF, OmpC and OmpA, respectively. SDS-PAGE analysis suggested that cell envelopes of staA or staA+ strains contained similar amounts of these proteins but characterization of strains carrying ompF (or C or A)-phoA gene fusions showed that mutation stA-2 reduced ompF gene expression by a factor of two. Analysis of double mutants strains carrying mutation staA-2 and a tolA, tolB, excC or excD periplasmic-leaky mutation showed that staA suppression was allele specific which suggested that proteins TolA and StaA might directly interact.  相似文献   

15.
Why does the human factor IX gene have a G + C content of 40%?   总被引:20,自引:2,他引:18       下载免费PDF全文
The factor IX gene has a G + C content of approximately 40% in all mammalian species examined. In human factor IX, C----T and G----A transitions at the dinucleotide CpG are elevated at least 24-fold relative to other transitions. Can the G + C content be explained solely by this hot spot of mutation? Using our mathematical model, we show that the elevation of mutation at CpG cannot alone lower the G + C content below 45%. To search for other hot spots of mutation that might contribute to the reduction of G + C content, we assessed the relative rates of base substitution in our sample of 160 families with hemophilia B. Seventeen independent single-base substitutions are reported herein for a total of 96 independent point mutations in our sample. The following conclusions emerge from the analysis of our data and, where appropriate, the data of others: (1) Transversions at CpG are elevated an estimated 7.7-fold relative to other transversions. (2) The mutation rates at non-CpG dinucleotides are remarkably uniform; none of the observed rates are either more than twofold above the median for transitions or more than threefold above the median for transversions. (3) The pattern of recent mutation is compatible with the pattern during mammalian evolution that has maintained the G + C content of the factor IX gene at approximately 40%.  相似文献   

16.
Interaction of negative (CytR) and positive (cAMP-CRP) control in the promoter region of the uridine phosphorylase (udp) gene of Escherichia coli has been studied by using udp-lac operon fusions in which the structural lacZ gene is expressed from the wild type promoter udpP+ or from mutant promoters udpP1 and udpP18. The specific activity of beta-galactosidase was examined in these fusions in cytR+ and cytR- backgrounds after introduction of specific mutations in crp locus, crp* and crp(a) altering interaction of CRP protein with catabolite-sensitive promoters. The data obtained using crp* mutation confirm the proposed model of the udp gene regulation, according to which CytR repressor protein interferes with CRP binding site in the promoter-operator region of the udp gene and thereby prevents the positive action of cAMP-CRP complex on the udp expression. Additional data in favor of this model were obtained using crp(a) mutation which most probably alters the structure of CRP protein in such a way that it exhibits more high affinity to the udp promoter, as compared to the CytR repressor protein. Indeed, taken by itself, the crp(a) mutation did not lead to any increase in the expression of udpP+-lac fusion under the conditions of cAMP limitation (on glucose-grown cells), in spite of whether or not the CytR repressor was present. However, when combined with the ptsG mutation or when cells were grown on succinate medium, complete constitutive expression of udpP+-lac fusion is observed, even in the presence of the cytR gene product. The effect of the crp(a) mutation was virtually the same in strains harboring udpP1-lac fusion. These data are in accordance with suggestion that udpP1 is a mutation in the site of the promoter-operator region that responds to the cytR gene product, while the corresponding binding site for CRP protein is still unaltered in this mutant. On the other hand, the crp(a) mutation causes only slight alteration in the expression of udpP18-lac fusion, providing additional evidence that udpP18 mutation seems to comprise a modification of the promoter-operator region, where binding sites for CRP and CytR proteins overlap.  相似文献   

17.
We have identified a spontaneous mitochondrial mutation, mfs-1 (mitochondrial frameshift suppressor-1), which suppresses a + 1 frameshift mutation localized in the yeast mitochondrial oxi1 gene. The suppressor strain exhibits a single base change (C to U) at position 42 of the mitochondrial serine-tRNA (UCN). To our knowledge, this is the first reported case showing that a mutation in the anticodon stem of a tRNA can cause frameshift suppression. The expression and aminoacylation of the mutant tRNASer(UCN) are not significantly affected. However, the base change at position 42 has two effects: first, residue U27 of the mutant tRNA is not modified to pseudouridine as observed in wild-type tRNASer(UCN). Second, the base change and/or the lack of modification of U27 leads to an alteration in the secondary/tertiary structure of the mutant tRNA. It is possible that there are such structural changes in the anticodon loop that enable the tRNA to read a four base codon, UCCA, thus restoring the wild-type reading frame.  相似文献   

18.
The majority of antibodies directed against p-azophenylarsonate (Ars) protein conjugates elicited during secondary immune responses of A/J mice bear a heritable cross-reactive Id (CRIa or IdCR) which corresponds to the utilization of a unique combination of variable region gene segments that can differ by somatic mutations. One such monoclonal anti-Ars antibody, 44-10, bears IdCR as defined by rabbit antisera but does not react with two anti-idiotypic mAb, 5Ci and AD8, which react with all primary (unmutated) IdCR+ antibodies and some secondary response IdCR+ antibodies. We therefore determined the complete sequence of antibody 44-10, which differs from the germline encoded (unmutated) IdCR+ antibody 36-65 at four positions in the H chain V region (VH): position 55 in the second complementarity determining region, 100 and 107 (D-gene junctions) and 110 (in JH2). The 44-10 L chain is unmutated. Sequence analyses of five other secondary immune response anti-Ars IdCR+ antibodies chosen on the basis of sharing one or more of the amino acid substitutions found in 44-10, were correlated with idiotypic expression of this set of antibodies. The results suggest that the mutation at VH position 55 (Asn----Lys) is responsible for loss of the 5Ci idiotope. To substantiate this hypothesis, oligonucleotide-directed mutagenesis of the germline encoded (unmutated) IdCR+ antibody was used to produce two mutants, one with VH Lys 55 and the other containing residues at positions 100, 107 and 110 identical to those found in 44-10. Id binding studies on these mutants confirm that 5Ci idiotope loss is due to conformational changes resulting from a mutation at VH position 55. This mutation also results in loss of the AD8 idiotope in the structural context of antibody 44-10.  相似文献   

19.
Transport and phosphorylation of glucose via enzymes II-A/II-B and II-BGlc of the phosphoenolpyruvate:sugar phosphotransferase system are tightly coupled in Salmonella typhimurium. Mutant strains (pts) that lack the phosphorylating proteins of this system, enzyme I and HPr, are unable to transport or to grow on glucose. From ptsHI deletion strains of S. typhimurium, mutants were isolated that regained growth on and transport of glucose. Several lines of evidence suggest that these Glc+ mutants have an altered enzyme II-BGlc as follows. (i) Insertion of a ptsG::Tn10 mutation (resulting in a defective II-BGlc) abolished growth on and transport of glucose in these Glc+ strains. Introduction of a ptsM mutation, on the other hand, which abolishes II-A/II-B activity, had no effect. (ii) Methyl alpha-glucoside transport and phosphorylation (specific for II-BGlc) was lowered or absent in ptsH+,I+ transductants of these Glc+ strains. Transport and phosphorylation of other phosphoenolpyurate:sugar phosphotransferase system sugars were normal. (iii) Membranes isolated from these Glc+ mutants were unable to catalyze transphosphorylation of methyl alpha-glucoside by glucose 6-phosphate, but transphosphorylation of mannose by glucose 6-phosphate was normal. (iv) The mutation was in the ptsG gene or closely linked to it. We conclude that the altered enzyme II-BGlc has acquired the capacity to transport glucose in the absence of phosphoenolpyruvate:sugar phosphotransferase system-mediated phosphorylation. However, the affinity for glucose decreased at least 1,000-fold as compared to the wild-type strain. At the same time the mutated enzyme II-BGlc lost the ability to catalyze the phosphorylation of its substrates via IIIGlc.  相似文献   

20.
Mutations in the gene for guanylate cyclase-activating protein-1 (GCAP1) (GUCA1A) have been associated with autosomal dominant cone dystrophy (COD3). In the present study, a severe disease phenotype in a large white family was initially shown to map to chromosome 6p21.1, the location of GUCA1A. Subsequent single-stranded conformation polymorphism analysis and direct sequencing revealed an A464G transition, causing an E155G substitution within the EF4 domain of GCAP1. Modeling of the protein structure shows that the mutation eliminates a bidentate amino acid side chain essential for Ca2+ binding. This represents the first disease-associated mutation in GCAP1, or any neuron-specific calcium-binding protein within an EF-hand domain, that directly coordinates Ca2+. The functional consequences of this substitution were investigated in an in vitro assay of retinal guanylate cyclase activation. The mutant protein activates the cyclase at low Ca2+ concentrations but fails to inactivate at high Ca2+ concentrations. The overall effect of this would be the constitutive activation of guanylate cyclase in photoreceptors, even at the high Ca2+ concentrations of the dark-adapted state, which may explain the dominant disease phenotype.  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