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 718 毫秒
1.
Theodore Homyk  Jr.  Angel Rodriguez    Jon Weil 《Genetics》1976,83(3):477-487
In the course of isolating viable T4 deletions that affect plaque morphology (Homyk and Weil 1974), two closely linked point mutants, sip1 and sip2, were obtained. They map between genes t and 52, cause a reduction in plaque size and burst size, and partially suppress the lethality of rII mutants for growth in lambda lysogens. These characteristics demonstrate that sip1 and sip2 are similar to mutants previously reported by Freedman and Brenner (1972). In addition, D. Hall (personal communication) has shown that sip1 and sip2 are similar to the mutant farP85, which affects the regulation of a number of early genes ( Chace and Hall 1975).——Sip suppression of rII mutants can be demonstrated in one-step growth experiments, even when both rII genes are completely deleted. This indicates that sip mutants do not simply reduce the level of rII gene products required for growth in a lambda lysogen. Instead, they alter the growth cycle so as to partially circumvent the need for any rII products.——Mutations at two other sites, designated L1 and L2, reverse the poor phage growth caused by sip and, in the one case tested, reverse the rII-suppressing ability of sip.  相似文献   

2.
As part of a study of protein folding, we have constructed a fine-structure map of 9 existing and 29 newly isolated UV- and hydroxylamine-induced temperature-sensitive (ts) mutations in gene 9 of Salmonella bacteriophage P22. Gene 9 specifies the polypeptide chain of the multimeric tail spikes, six of which form the cell attachment organelle of the phage. The 38 ts mutants were mapped against deletion lysogens with endpoints in gene 9. They mapped in 10 of the 15 deletion intervals. Two- and three-factor crosses between mutants within each interval indicated that at least 31 ts sites are represented among the 38 mutants. To determine the distribution of ts sites within the physical map, we identified the protein fragments from infection of su- hosts with 10 gene 9 amber mutants. Their molecular weights, ranging from 13,900 to 55,000 daltons, were combined with the genetic data to yield a composite map of gene 9. The 31 ts sites were distributed through most of the gene, but were most densely clustered in the central third.—None of the ts mutant pairs tested exhibited intragenic complementation. Studies of the defective phenotypes of the ts mutants (Goldenberg and King 1981; Smith and King 1981) revealed that most do not affect the thermostability of the mature protein, but instead prevent the folding or subunit assembly of the mutant chains synthesized at restrictive temperature. Thus, many of thes ts mutations identify sites in the polypeptide chain that are critical for the folding or maturation of the tail-spike protein.  相似文献   

3.
A Second Informational Suppressor, SUP-7 X, in CAENORHABDITIS ELEGANS   总被引:15,自引:14,他引:1  
More than 30 independent suppressor mutations have been obtained in the nematode C. elegans through reversion analysis of two unc-13 mutants. Many of the new isolates map to the region of the previously identified informational suppressor, sup-5 III (Waterston and Brenner 1978). Several of the other suppressor mutations map to the left half of the X-linkage group and define a second suppressor gene, sup-7 X. In tests against 40 mutations in six genes, the sup-7(st5) allele was found to suppress to a greater extent the same alleles acted on by sup-5(e1464). Like sup-5(e1464), sup-7(st5) acts on null alleles of the myosin heavy-chain gene unc-54 I (MacLeod et al. 1977; MacLeod, Waterston and Brenner 1977) and the putative paramyosin gene unc-15 I (Waterston et al. 1977). Chemical analysis of unc-15(e1214); sup-7(st5) animals show that paramyosin is restored to more than 30% of the wild-type level.—As was observed for sup-5(e1464), suppression by sup-7(st5) is dose dependent and is greater in animals grown at 15° than at 25°. However, associated with this increased suppression is a decreased viability of sup-7(st5) homozygotes. Reversion of the lethality has resulted in the isolation of deficiency mutations that complement st5 lethality, but lack suppressor function. These properties of sup-7(st5) suggest that it, like sup-5(e1464), is an informational suppressor of null alleles, and its reversion via deficiencies further narrows the possible explanations of its action.  相似文献   

4.
Donahue TF  Henry SA 《Genetics》1981,98(3):491-503
An extensive genetic analysis of inositol auxotrophic mutants of yeast is reported. The analysis includes newly isolated mutants, as well as those previously reported (Culbertson and Henry 1975). Approximately 70% of all inositol auxotrophs isolated are shown to be alleles of the ino1 locus, the structural gene for inositol-1-phosphate synthase, the major enzyme involved in inositol biosynthesis. Alleles of two other loci, ino2 and ino4, comprise 9% of total mutants, with the remainder representing unique loci or complementation groups. The ino1 locus was mapped by trisomic analysis with an n + 1 disomic strain constructed with complementing alleles at this locus. The ino1 locus is shown to be located between ura2 (11.1 cm) and cdc6 (21.8 cm) on chromosome X. An extended map of chromosome X of yeast is presented. Unlike most yeast loci, but similar to the his1 locus, the ino1 locus lacks allelic representatives that are suppressible by known suppressors. This finding suggests that premature termination of translation of the ino1 gene product may be incompatible with cell viability.  相似文献   

5.
E. Zouros  W. van Delden 《Genetics》1982,100(2):307-314
In a larval esterase of Drosophila mojavensis there are alleles whose products preferentially hydrolyze α-naphthyl esters, whereas the majority of the alleles hydrolyze preferentially β-naphthyl esters. In a collection of laboratory stocks α alleles have a frequency of 15%. Three different mobilities of α alleles were discovered, suggesting a polymorphism rather than a single mutation event. If substrate-preference polymorphisms are common among "multiple-substrate" enzymes (category II of Gillespie and Langley 1974), allozyme variation at these enzyme loci may well be maintained by balancing selection.  相似文献   

6.
Until recently, little was known of the genetic constitution of the heterochromatic segments of the major autosomes of Drosophila melanogaster . Our previous report described the genetic dissection of the proximal, heterochromatic region of chromosome 2 of Drosophila melanogaster by means of a series of overlapping deficiencies generated by the detachment of compound second autosomes (Hilliker and Holm 1975). Analysis of these deficiencies by inter se complementation, pseudo-dominance tests with proximal mutations and allelism tests with known deficiencies provided evidence for the existence of at least two loci between the centromere and the light locus in 2L and one locus in 2R between the rolled locus and the centromere. These data in conjunction with cytological observations demonstrated that light and rolled and three loci lying between them are located within the proximal heterochromatin of the second chromosome.——The present report describes the further analysis of this region through the induction with ethyl methanesulphonate (EMS) of recessive lethals allelic to the 2L and 2R proximal deficiencies associated with the detachment products. Analysis of the 118 EMS-induced recessive lethals and visible mutations recovered provided evidence for seven loci in the 2L heterochromatin and six loci in the 2R heterochromatin, with multiple alleles being obtained for most sites. Of these loci, one in 2L and two in 2R fall near the heterochromatic-euchromatic junctions of 2L and 2R respectively. None of the 113 EMS lethals behaved as a deficiency, implying that the heterochromatic loci uncovered in this study represent nonrepetitive cistrons. Thus functional genetic loci are found in heterochromatin, albeit at a very low density relative to euchromatin.  相似文献   

7.
Male-Specific Lethal Mutations of DROSOPHILA MELANOGASTER   总被引:6,自引:5,他引:1  
A total of 7,416 ethyl methanesulfonate (EMS)-treated second chromosomes and 6,212 EMS-treated third chromosomes were screened for sex-specific lethals. Four new recessive male-specific lethal mutations were recovered. When in homozygous condition, each of these mutations kills males during the late larval or early pupal stages, but has no detectable effect in females. One mutant, mlets, is a temperature sensitive allele of maleless, mle (Fukunaga, Tanaka and Oishi 1975), while the other three mutants identify two new loci: male-specific lethal-1 (msl-1) (two alleles) at map position 2-53.3 and male-specific lethal-2 (msl-2) at 2-9.0.——The male-specific lethality associated with these mutants is not related to the sex per se of the mutant flies, since sex-transforming genes fail to interact with these mutations. Moreover, the presence or absence of a Y chromosome in males or females has no influence on the male-specific lethal action of these mutations. Finally, no single region of the X chromosome, when present as a duplication, is sufficient to rescue males from the lethal effects of msl-1 or msl-2. These results suggest that the number of complete X chromosomes determines whether a fly homozygous for a male-specific lethal mutation lives or dies.  相似文献   

8.
9.
Homyk, Rodriguez and Weil (1976) have described T4 mutants, called sip, that partially suppress the inability of T4rII mutants to grow in λ lysogens. We have found that mutants sip1 and sip2 are resistant to folate analogs and overproduce FH2 reductase. The results of recombination and complementation studies indicate that sip mutations are in the mot gene. Like other mot mutations (Mattson, Richardson and Goodin 1974; Chace and Hall 1975; Sauerbier, Hercules and Hall 1976), the sip2 mutation affects the expression of many genes and appears to affect promoter utilization. The mot gene function is not required for T4 growth on most hosts, but we have found that it is required for good growth on E. coli CTr5X. Homyk, Rodriguez and Weil (1976) also described L mutations that reverse the effects of sip mutations. L2 decreases the folate analog resistance and the inability of sip2 to grow on CTr5X. L2 itself is partially resistant to a folate analog, and appears to reverse the effects of sip2 on gene expression. These results suggest that L2 affects another regulatory gene related to the mot gene.  相似文献   

10.
Zhang XS 《Genetics》2008,180(1):687-695
Why does phenotypic variation increase upon exposure of the population to environmental stresses or introduction of a major mutation? It has usually been interpreted as evidence of canalization (or robustness) of the wild-type genotype; but an alternative population genetic theory has been suggested by J. Hermisson and G. Wagner: “the release of hidden genetic variation is a generic property of models with epistasis or genotype–environment interaction.” In this note we expand their model to include a pleiotropic fitness effect and a direct effect on residual variance of mutant alleles. We show that both the genetic and environmental variances increase after the genetic or environmental change, but these increases could be very limited if there is strong pleiotropic selection. On the basis of more realistic selection models, our analysis lends further support to the genetic theory of Hermisson and Wagner as an interpretation of hidden variance.A common experimental observation in quantitative genetics is a higher phenotypic variance for quantitative traits in populations that carry a major mutation or are exposed to environmental stresses (e.g., heat shock) (Scharloo 1991; for a recent review see Gibson and Dworkin 2004). Part of the added variance must be genetic because the population responds to artificial selection. The lower variability of the wild type than that of the mutants has been interpreted as evidence for robustness or canalization (Waddington 1957): that is, under the new condition the magnitudes of gene effects across all trait loci increase relative to the original condition. The importance of canalization has been recognized for a long time and has been the subject of renewed interest recently (see de Visser et al. 2003 and Hansen 2006 for reviews).An alternative population genetic theory has been proposed by Hermisson and Wagner (2004), who suggest that the increase in genetic variance VG after the change in environmental conditions or genetic background is a generic property of the population, with no need to introduce canalization (Waddington 1957). The theory appears simple. Under mutation–selection balance (MSB), the mutant alleles are at a selective disadvantage and there is a negative correlation between frequencies and effects of mutations: mutant alleles of small effects on the trait segregate at intermediate frequencies. After the change in genetic or environmental background, gene effects consequently change due to G × E interaction or epistasis, which reduces the negative correlation because genes that were previously of small effects and at intermediate frequencies may now have large effects. That is, the frequencies of alleles are determined by the previous MSB, while their new effects are at least partly determined by the new conditions. The genetic variance will therefore increase.Hermisson and Wagner (2004) found that the predicted increase in genetic variance can be substantial; however, the predicted increase is highly sensitive to the population size and can increase without bound with increasing population size (see their Figure 2 and Equation 16). Genetic variance would enlarge with the population size within a small population (Lynch and Hill 1986; Weber and Diggins 1990), but becomes insensitive to the population size within large populations (Falconer and Mackay 1996, Chap. 20). Hence the unbounded increase under the novel environmental condition appears to us as a downside of their theory, even though the predicted increase can be reduced if the changed environmental condition is not novel but there is previous adaptation to it (see their Figure 3).Open in a separate windowFigure 2.—Influence of the pleiotropic effect (sp) on the increase of genetic variance ΔG in units of the interaction parameter ξ for a “typical” situation with strength of stabilizing selection ω2 = 0.1μ2, mutation rate λ = 0.1 per haploid genome per generation, and population size Ne = 106. The allelic pleiotropic effect on fitness and its variance effect on the trait independently follow gamma distributions with shape parameters βs and βv, respectively. The mean of a2 across loci is E(v) = E(a2) = 10−4μ2.Open in a separate windowOpen in a separate windowFigure 3.—Influence of shapes of distributions of mutational effects on (a) the variances at mutation–selection balance and (b) their increases after the genetic or environmental change. The squares represent the genetic variance and its increase and the triangles the environmental variance and its increase. The mutation rate is λ= 0.1 per haploid genome per generation, the population size is Ne = 109, and the strength of real stabilizing selection is ω2 = 0.1μ2. Allelic effects on trait value (a), fitness (s), and residual variance (b) are assumed to be independently distributed such that v = a2 follows a gamma () distribution with mean 10−4μ2, s follows gamma (βs) with mean sp = 0.05, and b follows gamma (βb) with mean 10−4μ2.The basic model that Hermisson and Wagner (2004) employed is that the quantitative trait is under real stabilizing selection and mutant alleles have effects on the focal trait only by changing its so-called locus genetic variance. At the mutation–real stabilizing selection balance, some mutants can segregate at intermediate frequencies because of their small effects and therefore weak selection; and there are more such mutants the more strongly leptokurtic is the distribution of effects at individual loci. The unbounded increase of Hermisson and Wagner (2004) results from such a gene-frequency distribution; but it has been shown (see Barton and Turelli 1989; Falconer and Mackay 1996; Lynch and Walsh 1998) that solely stabilizing selection, whether modeled with a Gaussian (Kimura 1965) or a house of-cards approximation (Turelli 1984) or even the generalized form of Hermisson and Wagner (2004) (i.e., their Equation 14), cannot provide a satisfactory explanation for the high levels of genetic variance observed in natural populations under realistic values of mutation and selection parameters.A common observation is that one trait is controlled by many genes and one gene can influence many traits; i.e., pleiotropy is ubiquitous (Barton and Turelli 1989; Barton and Keightley 2002; Mackay 2004; Ostrowski et al. 2005). Recent detailed studies suggest that pleiotropy calculated as the number of phenotypic traits affected varies considerably among quantitative trait loci (QTL) (Cooper et al. 2007; Albert et al. 2008; Kenney-Hunt et al. 2008; Wagner et al. 2008). Such pleiotropic effects must influence the magnitude of the variance. Though some genes have little effect on the focal trait, they almost certainly affect other traits and therefore are not neutral. The inclusion of pleiotropic effects on fitness strengthens the overall selection on mutant alleles and, assuming such pleiotropic effects are mainly deleterious, maintains them at low frequencies. The genetic variance for a trait is therefore likely to be maintained at lower levels than that under only real stabilizing selection on the trait alone (Tanaka 1996). Although the gene-frequency distribution is much more extreme under this joint model, the relevant rate of mutation is genomewide and hence is much larger than that where mutation affects only the focal trait as is assumed in the real stabilizing selection model (Turelli 1984; Falconer and Mackay 1996). Taking into account empirical knowledge of mutation parameters, a combination of both pleiotropic and real stabilizing selection appears to be a plausible mechanism for the maintenance of quantitative genetic variance (Zhang et al. 2004). If pleiotropic selection is much stronger than real stabilizing selection, the association between frequency and effect of mutant alleles is weaker than that for a real stabilizing selection model. Further, if overall selection is stronger than recurrent mutation, the frequency distribution of mutant alleles will be extreme. Under those situations, the increase of genetic variance after the genetic or environmental change will be kept at lower levels than that of Hermisson and Wagner (2004), and hence the unbounded increase could be avoided.Further, Hermisson and Wagner (2004) assume that the environmental variance is not under genetic control (i.e., the variance of phenotypic value given genotypic value is the same for all genotypes) and therefore is not subject to change. This assumption conflicts with the increasingly accumulating empirical data that indicate otherwise (Zhang and Hill 2005; Mulder et al. 2007 for reviews). Direct experimental evidence is available that mutation can directly affect environmental variance, VE (Whitlock and Fowler 1999; Mackay and Lyman 2005), and Baer (2008) provides what is perhaps the first clear demonstration that mutations increase environmental variances, on the basis of data for body size and productivity of Caenorhabditis elegans, and finds that the magnitudes of the increases are of the same order as those in the genetic variance.As real stabilizing selection on phenotype favors genotypes possessing low VE (Gavrilets and Hastings 1994; Zhang and Hill 2005), a mutant that contributes little to VE is more favored by stabilizing selection than one that contributes a lot. With all else being the same, mutants with small effect on VE thus segregate at relatively high frequencies at MSB. That is, there is a negative correlation between the effect on VE and the frequency of mutant genes. After the genetic or environmental change, some mutants that were previously of small effects on VE have large effects due to G × E interaction or epistasis while their frequencies remain roughly the same as in the previous MSB. This certainly increases environmental variance.In this note, we first assume that mutant alleles can affect only the mean value of a focal quantitative trait and otherwise affect fitness through their pleiotropic effects (Zhang et al. 2004) and try to answer the following questions: How will the conclusion of Hermisson and Wagner (2004) be affected by taking into account the pleiotropic effect of mutants? Can the “unbounded increase” be avoided? We then further assume that mutant alleles can also directly affect the environmental variance of the focal trait (Zhang and Hill 2008) and investigate how both VG and VE change following the genetic or environmental change in the population.  相似文献   

11.
The Origin of Spontaneous Mutation in SACCHAROMYCES CEREVISIAE   总被引:4,自引:2,他引:2  
Characterization of two antimutator loci in yeast shows that both are members of the same mutagenic repair system known to be responsible for almost all induced mutation (Lawrence and Christensen 1976, 1979a,b; Prakash 1976). One of the these newly isolated antimutator mutations is an allele of rev3 (Lemontt 1971b). Two other alleles of rev3 were tested and were also found to be antimutators. Double mutants carrying rev3 and mutator mutations of rad3, rad51 or rad18 are like rev3 single mutants with respect to spontaneous mutation rate, supporting the hypothesis (Hastings, Quah and von Borstel 1976) that many mutators in yeast act by channelling spontaneous lesions from accurate to mutagenic repair. However, the enhanced mutation rate seen in a radiation-resistant mutator mutant mut1 is not dependent on REV3, but is dependent on another gene designated ANT1. An additive effect on the reduction in spontaneous mutation, seen in the ant1 rev3 double-mutant strain, leads to the conclusion that at least 90% of spontaneous mutations seen in the wild type are caused by mutagenic repair of spontaneous lesions.  相似文献   

12.
Based on the Fisher-Muller theory of the evolution of recombination, an argument can be constructed predicting that a recessive allele favoring recombination will be favored, if there are either favorable or deleterious mutants occurring at other loci. In this case there is no clear distinction between individual and group selection. Computer simulation of populations segregating for recessive or dominant recombination alleles showed selection favoring recombination, except in the case of a dominant recombination allele with deleterious background mutants. The relationship of this work to parallel investigations by Williams and by Strobeck, Maynard Smith, and Charlesworth is explored. All seem to rely on the same phenomenon. There seems no reason to assume that the evolution of recombination must have occurred by group selection.  相似文献   

13.
14.
Sydney Brenner 《Genetics》2009,182(2):413-415
The replicative life span (RLS) of Saccharomyces cerevisiae has been established as a model for the genetic regulation of longevity despite the inherent difficulty of the RLS assay, which requires separation of mother and daughter cells by micromanipulation after every division. Here we present the mother enrichment program (MEP), an inducible genetic system in which mother cells maintain a normal RLS—a median of 36 generations in the diploid MEP strain—while the proliferative potential of daughter cells is eliminated. Thus, the viability of a population over time becomes a function of RLS, and it displays features of a survival curve such as changes in hazard rate with age. We show that viability of mother cells in liquid culture is regulated by SIR2 and FOB1, two opposing regulators of RLS in yeast. We demonstrate that viability curves of these short- and long-lived strains can be easily distinguished from wild type, using a colony formation assay. This provides a simplified screening method for identifying genetic or environmental factors that regulate RLS. Additionally, the MEP can provide a cohort of cells at any stage of their life span for the analysis of age-associated phenotypes. These capabilities effectively remove the hurdles presented by RLS analysis that have hindered S. cerevisiae aging studies since their inception 50 years ago.THE budding yeast Saccharomyces cerevisiae is a popular model system for studying fundamental processes of cellular aging (reviewed in Steinkraus et al. 2008). Analyses over the past 50 years have led to the idea that budding yeast can be used to study three types of cellular aging. Replicative aging describes the division potential of individual cells and relies on the asymmetric cell divisions of budding yeast that yield distinct mother and daughter cells. Replicative life span (RLS) is defined as the number of times an individual cell divides before it undergoes senescence (Mortimer and Johnston 1959). Chronological aging describes the capacity of cells in stationary phase (analogous to G0 in higher eukaryotes) to maintain viability over time, which is assayed by their ability to reenter the cell cycle when nutrients are reintroduced (Longo et al. 1996). Finally, budding yeast have been used to study clonal senescence, which is analogous to the Hayflick limit imposed on mammalian tissue culture cells and characterized by a finite number of times a population of cells can divide. Although wild-type yeast populations do not senesce, this phenomenon has been observed in mutant strains such as those lacking telomerase components (Lundblad and Szostak 1989; Singer and Gottschling 1994).While genetic screens have been applied to examine clonal and chronological aging (Lundblad and Szostak 1989; Powers et al. 2006; Murakami et al. 2008), they have been limited in their application to studying replicative aging (Kaeberlein and Kennedy 2005; Kaeberlein et al. 2005b). This limitation arises from the arduous nature of isolating replicatively aged yeast cells. The current “gold standard” for isolating aged mother cells is by micromanipulation, where daughter cells are counted and removed after every division (Park et al. 2002). Although micromanipulation is currently the only method capable of accurately measuring RLS in yeast, it is severely constrained by the small number of cells that can be analyzed. Thus, genetic analysis of the regulation of RLS has been limited to a candidate gene approach (reviewed in Steinkraus et al. 2008).True genetic analysis of RLS will require large populations of aged cells. However, there are two confounding issues that make isolation of aged individuals difficult. First, single-cell pedigree analysis has shown that age-associated phenotypes, such as replicative life span potential, segregate asymmetrically between mother and daughter cells, rendering age-associated phenotypes nonheritable (Egilmez and Jazwinski 1989; Kennedy et al. 1994). Thus, daughter cells are generally “reset” to a young state with every generation. Second, when age is measured in terms of cell divisions, an unfractionated population is predominately young. The fraction of the population at an age of n cell divisions is ∼1/2n. Individual cells that reach the median RLS, which is ∼26 generations for haploid cells of the S288C strain background (Kaeberlein et al. 2005a), represent an insignificant fraction of the total population. In fact, it is unlikely that any cell reaches such an advanced age because nutrient depletion will limit the division potential of the population (Dickinson and Schweizer 1999).As an alternative to micromanipulation, methods were developed to isolate aged cells from liquid cultures (Smeal et al. 1996; Sinclair and Guarente 1997; Chen and Contreras 2007). However, due to the exponential growth of progeny cells, these populations are technically limited to 7–12 generations before nutrient depletion interferes with replicative aging. While sequential rounds of growth and purification are possible, the inability to continuously follow an undisturbed cohort of cells prevents the measurement of RLS by these methods. Instead, purification methods are primarily used for the examination of molecular changes associated with aging cells. Unfortunately, low yields and loss of viability due to purification methods diminish their utility for analyzing phenotypes that affect cells of advanced age. As an alternative to purification from natural populations, a strategy to genetically regulate the replicative capacity of daughter cells and avoid the limits imposed by exponential growth has been described (Jarolim et al. 2004). While this system effectively prevents division of daughter cells, it unintentionally decreases the median RLS of mother cells to four cell divisions, thus restricting its usefulness.Here we describe the development of a novel genetic selection against newborn daughter cells, the “mother enrichment program” (MEP), which restricts the replicative capacity of daughter cells while allowing mother cells to achieve a normal RLS. We demonstrate that upon induction of the selection, the viability of MEP strains growing in liquid culture is determined by the RLS of the initial population of mother cells. MEP cultures therefore allow the comparison of RLS between strains without the need for micromanipulation. Additionally, because MEP cultures are not subject to nutrient limitation, single-step affinity purification of aged cells can be achieved at any point during their life span. Together, these capabilities substantially resolve the technical hurdles that have made replicative aging studies in S. cerevisiae exceptionally challenging.  相似文献   

15.
The close linkage of the glnA gene with polA was exploited to construct a fine structure map of polA by means of generalized transduction with phage P1. Nine different polA- alleles were mapped by recombinational crosses. The results indicate a gene order consistent with previous observations (Kelley and Grindley 1976a; Murray and Kelley 1979). Three mutations, polA5, polA6 and polA12 map within the "carboxy-terminal" or "large-fragment" portion of the gene in unambiguous order. Four alleles, known to affect the "aminoterminal" portion of the gene, polA107, polA214, polA480ex and polA4113, appear to be closely linked with certain ambiguities in their exact order. All four of these mutations are known to alter the 5''→3'' exonuclease activity of DNA polymerase I and three of them result in the conditional lethal polA- phenotype. The polA1 nonsense mutation maps between these two groups in a position consistent with its known effect, production of an amber fragment that includes the 5''→3'' exonuclease. The final allele, resA1, is another nonsense mutation that maps at the extreme "amino-terminus" of the cistron.——A number of control experiments were conducted to determine the effects of polA- mutations on the P1-mediated recombinational event. These experiments indicated that abortive transduction occurs quite frequently, but the formation of abortive transductants and segregation of unselected transduced markers among daughter progeny is like that observed by other investigators. There was no evidence that any individual polA- allele behaved in an exceptional fashion during recombination.  相似文献   

16.
With the most recent releases of the Drosophila melanogaster genome sequences, much of the previously absent heterochromatic sequences have now been annotated. We undertook an extensive genetic analysis of existing lethal mutations, as well as molecular mapping and sequence analysis (using a candidate gene approach) to identify as many essential genes as possible in the centromeric heterochromatin on the right arm of the second chromosome (2Rh) of D. melanogaster. We also utilized available RNA interference lines to knock down the expression of genes in 2Rh as another approach to identifying essential genes. In total, we verified the existence of eight novel essential loci in 2Rh: CG17665, CG17683, CG17684, CG17883, CG40127, CG41265, CG42595, and Atf6. Two of these essential loci, CG41265 and CG42595, are synonymous with the previously characterized loci l(2)41Ab and unextended, respectively. The genetic and molecular analysis of the previously reported locus, l(2)41Ae, revealed that this is not a single locus, but rather it is a large region of 2Rh that extends from unextended (CG42595) to CG17665 and includes four of the novel loci uncovered here.THE term “heterochromatin” was introduced by Heitz (1928) to describe regions of mitotic chromosomes that remain condensed throughout the cell cycle, in contrast to regions of euchromatin, which condense only during cell division. Heterochromatin was later divided into two classes: constitutive and facultative heterochromatin (Brown 1966). Constitutive heterochromatin is found in large blocks near centromeres and telomeres, while facultative heterochromatin can be described as silenced euchromatin that undergoes heterochromatization at specific developmental stages. Other properties of constitutive heterochromatin include late replication in S phase, low gene density, strikingly reduced level of meiotic recombination, enrichment in transposable element sequences and highly repetitive satellite DNA sequences, and the ability to silence euchromatic gene expression in a phenomenon called position effect variegation.Approximately 30% of the Drosophila melanogaster genome consists of constitutive heterochromatin (Gatti and Pimpinelli 1992). Centromeric heterochromatin in D. melanogaster is composed of mainly middle-repeat satellite DNA sequences and clusters of transposable element sequences (Lohe et al. 1993; Pimpinelli et al. 1995). Genes that reside in the heterochromatin are scattered like islands between the satellites and clusters of transposable elements. On average, heterochromatic genes are larger than euchromatic genes, primarily due to the prevalent accumulation of transposable element sequences in their introns (Devlin et al. 1990; Biggs et al. 1994; Dimitri et al. 2003a,b; Hoskins et al. 2007). Heterochromatic genes also tend to be AT-rich compared to their euchromatic counterparts; there is some evidence suggesting that the coding sequences of heterochromatic genes evolve toward AT richness in response to being located in heterochromatin (Yasuhara et al. 2005; Díaz-Castillo and Golic 2007).Drosophila heterochromatin is vastly under-replicated in polytene chromosomes, so heterochromatic genes cannot easily be mapped through polytene analysis. However, by using Hoechst 33258 and N-chromosome banding techniques, Dimitri (1991) was successful in dividing heterochromatin in mitotic chromosomes into distinct cytological bands; this was an important step in mapping the precise location of heterochromatic genes because before this time heterochromatic genes could be mapped only relative to one another. Here we focus on further refining the previous mapping work on essential genes in the proximal heterochromatin of the right arm of the second chromosome (2Rh) in cytological region h41–h46 of D. melanogaster (Hilliker 1976; Hilliker et al. 1980; Coulthard et al. 2003; Myster et al. 2004).Early mapping studies in D. melanogaster putatively placed the light (lt) and rolled (rl) genes in, or near, chromosome 2 heterochromatin (Schultz 1936; Hannah 1951; Hessler 1958). The first large-scale mutagenesis specifically directed at finding vital loci in second chromosome heterochromatin was conducted by Hilliker (1976). Using heterochromatic deletions created by Hilliker and Holm (1975), Hilliker (1976) set out to map vital loci using the mutagen ethyl methanesulfonate (EMS). He identified seven individual lethal complementation groups in 2Rh that were interpreted as representing seven vital loci. One of these heterochromatic loci was identified as the previously described rl gene. Two of the remaining vital loci have since been identified: Nipped-A is synonymous with the l(2) 41Ah complementation group (Rollins et al. 1999) and RpL38 is synonymous with Minute(2)41A and Hilliker''s (1976) l(2)41Af complementation group (Marygold et al. 2005; also referred to as l(2)Ag in FlyBase). In addition, Rollins et al. (1999) found the Nipped-B gene to be located in 2Rh, but how this locus fit into the data from Hilliker (1976) was unclear.With the limited release of some of the more distal heterochromatic sequences (Hoskins et al. 2002), a more recent mutagenesis screen focusing on distal 2Rh was conducted by Myster et al. (2004). In the region defined by the overlap between Df(2R)41A8 and Df(2R)41A10 (the latter was previously shown to be deficient for most of 2Rh; Hilliker and Holm 1975), Myster et al. (2004) reported the existence of 15 vital loci, considerably more than the 4 essential loci predicted by Hilliker (1976). The discrepancy between these two studies was the catalyst for this current work. Each group used the same mutagen, EMS, yet each group came up with very different interpretations of the number of vital loci.Hilliker''s interpretation relied on earlier evidence that EMS preferentially produced point mutations and not large-scale aberrations (Lim and Snyder 1974). Assuming that the mutants isolated in his study were point mutations, or small aberrations limited to one locus, Hilliker found that some of the loci that he identified exhibited complex interallelic complementation; the most complex complementation pattern was observed with locus l(2)41Ae. On the other hand, the interpretation of Myster et al. (2004) was that heterochromatin was more sensitive to EMS and that EMS could produce large heterochromatic deletions; they proposed that the complex interallelic complementation in l(2)41Ae was due to the presence of deletions and that l(2)41Ae represented a region of 2Rh containing many genes, rather than being a single locus.To resolve these different interpretations of the genomic segment containing l(2)41Ae (i.e., is it a single locus or a region of 2Rh), we set out to map l(2)41Ae and the region surrounding the presumed location of l(2)41Ae (as in Myster et al. 2004) by performing a large-scale inter se complementation analysis between all available mutant lines that were previously mapped to l(2)41Ae (including Nipped-B). In addition, we undertook a molecular mapping and sequence analysis, using a candidate gene approach with the most recent annotation of 2Rh (Hoskins et al. 2007), to characterize the region and identify as many essential genes as possible. We also used these approaches to map l(2)41Ab and unextended [two of the more proximal complementation groups identified by Hilliker (1976)]. Finally, we also utilized available RNA interference (RNAi) lines to knock down the expression of 12 genes in 2Rh in an attempt to identify essential genes.  相似文献   

17.
Peptidoglycan hydrolases (PGHs) are responsible for bacterial cell lysis. Most PGHs have a modular structure comprising a catalytic domain and a cell wall-binding domain (CWBD). PGHs of bacteriophage origin, called endolysins, are involved in bacterial lysis at the end of the infection cycle. We have characterized two endolysins, Lc-Lys and Lc-Lys-2, identified in prophages present in the genome of Lactobacillus casei BL23. These two enzymes have different catalytic domains but similar putative C-terminal CWBDs. By analyzing purified peptidoglycan (PG) degradation products, we showed that Lc-Lys is an N-acetylmuramoyl-l-alanine amidase, whereas Lc-Lys-2 is a γ-d-glutamyl-l-lysyl endopeptidase. Remarkably, both lysins were able to lyse only Gram-positive bacterial strains that possess PG with d-Ala4d-Asx-l-Lys3 in their cross-bridge, such as Lactococcus casei, Lactococcus lactis, and Enterococcus faecium. By testing a panel of L. lactis cell wall mutants, we observed that Lc-Lys and Lc-Lys-2 were not able to lyse mutants with a modified PG cross-bridge, constituting d-Ala4l-Ala-(l-Ala/l-Ser)-l-Lys3; moreover, they do not lyse the L. lactis mutant containing only the nonamidated d-Asp cross-bridge, i.e. d-Ala4d-Asp-l-Lys3. In contrast, Lc-Lys could lyse the ampicillin-resistant E. faecium mutant with 3→3 l-Lys3-d-Asn-l-Lys3 bridges replacing the wild-type 4→3 d-Ala4-d-Asn-l-Lys3 bridges. We showed that the C-terminal CWBD of Lc-Lys binds PG containing mainly d-Asn but not PG with only the nonamidated d-Asp-containing cross-bridge, indicating that the CWBD confers to Lc-Lys its narrow specificity. In conclusion, the CWBD characterized in this study is a novel type of PG-binding domain targeting specifically the d-Asn interpeptide bridge of PG.  相似文献   

18.
The transport of some sugars at the antiluminal face of renal cells was studied using teased tubules of flounder (Pseudopleuronectes americanus). The analytical procedure allowed the determination of both free and total (free plus phosphorylated) tissue sugars. The inulin space of the preparation was 0.333 ± 0.017 kg/kg wet wt (7 animals, 33 analyses). The nonmetabolizable α-methyl-D-glucoside entered the cells by a carrier-mediated (phloridzin-sensitive), ouabain-insensitive process. The steady-state tissue/medium ratio was systematically below that for diffusion equilibrium. D-Glucose was a poor inhibitor of α-methyl-glucoside transport, D-galactose was ineffective. The phloridzin-sensitive transport processes of 2-deoxy-D-glucose,D-galactose,and 2-deoxy-D-galactose were associated with considerable phosphorylation. Kinetic evidence suggested that these sugars were transported in free form and subsequently were phosphorylated. 2-Deoxy-D-glucose accumulated in the cells against a slight concentration gradient. This transport was greatly inhibited by D-glucose, whereas α-methyl-glucoside and also D-galactose and its 2-deoxy-derivative were ineffective. D-Galactose and 2-deoxy-D-galactose mutually competed for transport; D-glucose, 2-deoxy-D-glucose, and α-methyl-D-glucoside were ineffective. Studies using various sugars as inhibitors suggest the presence of three carrier-mediated pathways of sugar transport at the antiluminal cell face of the flounder renal tubule: the pathway of α-methyl-D-glucoside (not shared by D-glucose); the pathway commonly shared by 2-deoxy-D-glucose and D-glucose; the pathway shared by D-galactose and 2-deoxy-D-galactose.  相似文献   

19.
The pathway of d-xylose degradation in archaea is unknown. In a previous study we identified in Haloarcula marismortui the first enzyme of xylose degradation, an inducible xylose dehydrogenase (Johnsen, U., and Schönheit, P. (2004) J. Bacteriol. 186, 6198–6207). Here we report a comprehensive study of the complete d-xylose degradation pathway in the halophilic archaeon Haloferax volcanii. The analyses include the following: (i) identification of the degradation pathway in vivo following 13C-labeling patterns of proteinogenic amino acids after growth on [13C]xylose; (ii) identification of xylose-induced genes by DNA microarray experiments; (iii) characterization of enzymes; and (iv) construction of in-frame deletion mutants and their functional analyses in growth experiments. Together, the data indicate that d-xylose is oxidized exclusively to the tricarboxylic acid cycle intermediate α-ketoglutarate, involving d-xylose dehydrogenase (HVO_B0028), a novel xylonate dehydratase (HVO_B0038A), 2-keto-3-deoxyxylonate dehydratase (HVO_B0027), and α-ketoglutarate semialdehyde dehydrogenase (HVO_B0039). The functional involvement of these enzymes in xylose degradation was proven by growth studies of the corresponding in-frame deletion mutants, which all lost the ability to grow on d-xylose, but growth on glucose was not significantly affected. This is the first report of an archaeal d-xylose degradation pathway that differs from the classical d-xylose pathway in most bacteria involving the formation of xylulose 5-phosphate as an intermediate. However, the pathway shows similarities to proposed oxidative pentose degradation pathways to α-ketoglutarate in few bacteria, e.g. Azospirillum brasilense and Caulobacter crescentus, and in the archaeon Sulfolobus solfataricus.d-Xylose, a constituent of the polymer xylan, is the major component of the hemicellulose plant cell wall material and thus one of the most abundant carbohydrates in nature. The utilization of d-xylose by microorganisms has been described in detail in bacteria and fungi, for which two different catabolic pathways have been reported. In many bacteria, such as Escherichia coli, Bacillus, and Lactobacillus species, xylose is converted by the activities of xylose isomerase and xylulose kinase to xylulose 5-phosphate as an intermediate, which is further degraded mainly by the pentose phosphate cycle or phosphoketolase pathway. Most fungi convert xylose to xylulose 5-phosphate via xylose reductase, xylitol dehydrogenase, and xylulose kinase. Xylulose 5-phosphate is also an intermediate of the most common l-arabinose degradation pathway in bacteria, e.g. of E. coli, via activities of isomerase, kinase, and epimerase (1).Recently, by genetic evidence, a third pathway of xylose degradation was proposed for the bacterium Caulobacter crescentus, in analogy to an alternative catabolic pathway of l-arabinose, reported for some bacteria, including species of Azospirillum, Pseudomonas, Rhizobium, Burkholderia, and Herbasprillum (2, 3). In these organisms l-arabinose is oxidatively degraded to α-ketoglutarate, an intermediate of the tricarboxylic acid cycle, via the activities of l-arabinose dehydrogenase, l-arabinolactonase, and two successive dehydration reactions forming 2-keto-3-deoxy-l-arabinoate and α-ketoglutarate semialdehyde; the latter compound is further oxidized to α-ketoglutarate via NADP+-specific α-ketoglutarate semialdehyde dehydrogenase (KGSADH).2 In a few Pseudomonas and Rhizobium species, a variant of this l-arabinose pathway was described involving aldolase cleavage of the intermediate 2-keto-3-deoxy-l-arabinoate to pyruvate and glycolaldehyde, rather than its dehydration and oxidation to α-ketoglutarate (4). Because of the presence of some analogous enzyme activities in xylose-grown cells of Azosprillum and Rhizobium, the oxidative pathway and its variant was also proposed as a catabolic pathway for d-xylose. Recent genetic analysis of Caulobacter crecentus indicates the presence of an oxidative pathway for d-xylose degradation to α-ketoglutarate. All genes encoding xylose dehydrogenase and putative lactonase, xylonate dehydratase, 2-keto-3-deoxylonate dehydratase, and KGSADH were found to be located on a xylose-inducible operon (5). With exception of xylose dehydrogenase, which has been partially characterized, the other postulated enzymes of the pathway have not been biochemically analyzed.The pathway of d-xylose degradation in the domain of archaea has not been studied so far. First analyses with the halophilic archaeon Haloarcula marismortui indicate that the initial step of d-xylose degradation involves a xylose-inducible xylose dehydrogenase (6) suggesting an oxidative pathway of xylose degradation to α-ketoglutarate, or to pyruvate and glycolaldehyde, in analogy to the proposed oxidative bacterial pentose degradation pathways. Recently, a detailed study of d-arabinose catabolism in the thermoacidophilic crenarchaeon Sulfolobus solfataricus was reported. d-Arabinose was found to be oxidized to α-ketoglutarate involving d-arabinose dehydrogenase, d-arabinoate dehydratase, 2-keto-3-deoxy-d-arabinoate dehydratase, and α-ketoglutarate semialdehyde dehydrogenase (3).In this study, we present a comprehensive analysis of the complete d-xylose degradation pathway in the halophilic archaeon Haloferax volcanii. This halophilic archaeon was chosen because it exerts several suitable properties for the analyses. For example, it can be cultivated on synthetic media with sugars, e.g. xylose, an advantage for in vivo labeling studies in growing cultures. Furthermore, a shotgun DNA microarray of H. volcanii is available (7) allowing the identification of xylose-inducible genes, and H. volcanii is one of the few archaea for which an efficient protocol was recently described to generate in-frame deletion mutants.Accordingly, the d-xylose degradation pathway was elucidated following in vivo labeling experiments with [13C]xylose, DNA microarray analyses, and the characterization of enzymes involved and their encoding genes. The functional involvement of genes and enzymes was proven by constructing corresponding in-frame deletion mutants and their analysis by selective growth experiments on xylose versus glucose. The data show that d-xylose was exclusively degraded to α-ketoglutarate involving xylose dehydrogenase, a novel xylonate dehydratase, 2-keto-3-deoxyxylonate dehydratase, and α-ketoglutarate semialdehyde dehydrogenase.  相似文献   

20.
Thermotoga maritima is a Gram-negative, hyperthermophilic bacterium whose peptidoglycan contains comparable amounts of l- and d-lysine. We have determined the fine structure of this cell-wall polymer. The muropeptides resulting from the digestion of peptidoglycan by mutanolysin were separated by high-performance liquid chromatography and identified by amino acid analysis after acid hydrolysis, dinitrophenylation, enzymatic determination of the configuration of the chiral amino acids, and mass spectrometry. The high-performance liquid chromatography profile contained four main peaks, two monomers, and two dimers, plus a few minor peaks corresponding to anhydro forms. The first monomer was the d-lysine-containing disaccharide-tripeptide in which the d-Glu-d-Lys bond had the unusual γ→ϵ arrangement (GlcNAc-MurNAc-l-Ala-γ-d-Glu-ϵ-d-Lys). The second monomer was the conventional disaccharide-tetrapeptide (GlcNAc-MurNAc-l-Ala-γ-d-Glu-l-Lys-d-Ala). The first dimer contained a disaccharide-l-Ala as the acyl donor cross-linked to the α-amine of d-Lys in a tripeptide acceptor stem with the sequence of the first monomer. In the second dimer, donor and acceptor stems with the sequences of the second and first monomers, respectively, were connected by a d-Ala4-α-d-Lys3 cross-link. The cross-linking index was 10 with an average chain length of 30 disaccharide units. The structure of the peptidoglycan of T. maritima revealed for the first time the key role of d-Lys in peptidoglycan synthesis, both as a surrogate of l-Lys or meso-diaminopimelic acid at the third position of peptide stems and in the formation of novel cross-links of the l-Ala1(α→α)d-Lys3 and d-Ala4(α→α)d-Lys3 types.Peptidoglycan (or murein) is a giant macromolecule whose main function is the protection of the cytoplasmic membrane against the internal osmotic pressure. It is composed of alternating residues of N-acetylglucosamine (GlcNAc) and N-acetylmuramic acid (MurNAc)2 cross-linked by short peptides (1). The composition of the peptide stem in nascent peptidoglycan is l-Ala1-γ-d-Glu2-X3-d-Ala4-d-Ala5, where X is most often meso-diaminopimelic acid (meso-A2pm) or l-lysine in Gram-negative and Gram-positive species, respectively (2, 3). In the mature macromolecule, the last d-Ala residue is removed. Cross-linking of the glycan chains generally occurs between the carboxyl group of d-Ala at position 4 of a donor peptide stem and the side-chain amino group of the diamino acid at position 3 of an acceptor peptide stem (4→3 cross-links). Cross-linking is either direct or through a short peptide bridge such as pentaglycine in Staphylococcus aureus (2, 3). The enzymes for the formation of the 4→3 cross-links are active-site serine dd- transpeptidases that belong to the penicillin-binding protein (PBP) family and are the essential targets of β-lactam antibiotics in pathogenic bacteria (4). Catalysis involves the cleavage of the d-Ala4-d-Ala5 bond of a donor peptide stem and the formation of an amide bond between the carboxyl of d-Ala4 and the side chain amine at the third position of an acceptor stem. Transpeptidases of the ld specificity are active-site cysteine enzymes that were shown to act as surrogates of the PBPs in mutants of Enterococcus faecium resistant to β-lactam antibiotics (5). They cleave the X3-d-Ala4 bond of a donor stem peptide to form 3→3 cross-links. This alternate mode of cross-linking is usually marginal, although it has recently been shown to predominate in non-replicative “dormant” forms of Mycobacterium tuberculosis (6).Thermotoga maritima is a Gram-negative, extremely thermophilic bacterium isolated from geothermally heated sea floors by Huber et al. (7). A morphological characteristic is the presence of an outer sheath-like envelope called “toga.” Although the organism has received considerable attention for its biotechnological potential, studies about its peptidoglycan are scarce (811), and in particular the fine structure of the macromolecule is still unknown. In their initial work, Huber et al. (7) showed that the composition of its peptidoglycan was unusual for a Gram-negative species, because it contained both isomers of lysine and no A2pm. Recently, we purified and studied the properties of T. maritima MurE (12); this enzyme is responsible for the addition of the amino acid residue at position 3 of the peptide stem (13, 14). We demonstrated that T. maritima MurE added in vitro l- and d-Lys to UDP-MurNAc-l-Ala-d-Glu. Although l-Lys was added in the usual way, yielding the conventional nucleotide UDP-MurNAc-l-Ala-γ-d-Glu-l-Lys containing a d-Glu(γ→α)l-Lys amide bond, the d-isomer was added in an “upside-down” manner, yielding the novel nucleotide UDP-MurNAc-l-Ala-d-Glu(γ→ϵ)d-Lys. We also showed that the d-Lys-containing nucleotide was not a substrate for T. maritima MurF, the subsequent enzyme in the biosynthetic pathway, whereas this ligase catalyzed the addition of dipeptide d-Ala-d-Ala to the l-Lys-containing tripeptide, yielding the conventional UDP-MurNAc-pentapeptide (12).However, both the l-Lys-containing UDP-MurNAc-pentapeptide and d-Lys-containing UDP-MurNAc-tripeptide were used as substrates by T. maritima MraY with comparable efficiencies in vitro (12). This observation implies that the unusual d-Lys-containing peptide stems are likely to be translocated to the periplasmic face of the cytoplasmic membrane and to participate in peptidoglycan polymerization. Therefore, we have determined here the fine structure of T. maritima peptidoglycan and we have shown that l-Lys- and d-Lys-containing peptide stems are both present in the polymer, the latter being involved in the formation of two novel types of peptidoglycan cross-link.  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