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 46 毫秒
1.
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   

2.
3.
4.
5.
6.
7.
Mlh1p forms three heterodimers that are important for mismatch repair (Mlh1p/Pms1p), crossing over during meiosis (Mlh1p/Mlh3p), and channeling crossover events into a specific pathway (Mlh1p/Mlh2p). All four proteins contain highly conserved ATPase domains and Pms1p has endonuclease activity. Studies of the functional requirements for Mlh1p/Pms1p in Saccharomyces cerevisae revealed an asymmetric contribution of the ATPase domains to repairing mismatches. Here we investigate the functional requirements of the Mlh1p and Mlh3p ATPase domains in meiosis by constructing separation of function mutations in Mlh3p. These mutations are analogous to mutations of Mlh1p that have been shown to lead to loss of ATP binding and/or ATP hydrolysis. Our data suggest that ATP binding by Mlh3p is required for meiotic crossing over while ATP hydrolysis is dispensable. This has been seen previously for Mlh1p. However, when mutations that affect ATP hydrolysis by both Mlh3p and Mlh1p are combined within a single cell, meiotic crossover frequencies are reduced. These observations suggest that the function of the Mlh1p/Mlh3p heterodimer requires both subunits to bind ATP but only one to efficiently hydrolyze it. Additionally, two different amino acid substitutions to the same residue (G97) in Mlh3p affect the minor mismatch repair function of Mlh3p while only one of them compromises its ability to promote crossing over. These studies thus reveal different functional requirements among the heterodimers formed by Mlh1p.CROSSING over during meiosis not only generates variation but is also important for providing the necessary interactions between homologous chromosomes that ensure correct segregation at division I of meiosis. Recombination is initiated by the production of programmed double-strand breaks (DSBs), catalyzed by the covalently attached Spo11p (Bergerat et al. 1997; Keeney et al. 1997), aided by a number of proteins (reviewed in Keeney and Neale 2006). DSBs are made at a much higher frequency than crossovers, and designation of only a subset to yield crossovers is thought to occur during early stages of DSB repair (Borner et al. 2004). At least two distinct pathways contribute to the production of crossover events in Saccharomyces cerevisiae. The major pathway is dependent on Msh4p/Msh5p and the mismatch repair proteins Mlh1p and Mlh3p (Ross-MacDonald and Roeder 1994; Hollingsworth et al. 1995; Hunter and Borts 1997; Wang et al. 1999; Abdullah et al. 2004) and the second pathway is dependent on Mus81p/Mms4p endonuclease (de los Santos et al. 2001, 2003).Mitotic mismatch repair (MMR) is the process by which mutations that arise during DNA replication and recombination are recognized and removed (reviewed in Kolodner 1996; Harfe and Jinks-Robertson 2000). Msh2p forms a heterodimer with Msh6p (MutSα) to repair base–base mismatches and small insertions and/or deletions and with Msh3p (MutSβ) to repair large insertions and/or deletions (reviewed in Jiricny 2006). Mlh1p forms heterodimers with Pms1p, Mlh2p, and Mlh3p to coordinate the removal of these mismatches (Prolla et al. 1994; Wang et al. 1999). Mlh1p/Pms1p (MutLα) are involved in the repair of all types of mismatches in combination with MutSα and MutSβ, and in the absence of either protein a mutator phenotype is observed (Habraken et al. 1997, 1998). Mlh1p/Mlh2p (MutLβ) and Mlh1p/Mlh3p (MutLγ) are involved in the MutSβ pathway only, which repairs frameshift mutations caused by insertions or deletions. Consequently mlh3Δ mutants only exhibit a weak mutator phenotype, due to a lesser involvement in mismatch repair and a partial overlap in function with Pms1p (Flores-Rozas and Kolodner 1998; Harfe et al. 2000).Although the MutL homologs interact primarily through their C-terminal domains (Pang et al. 1997; Ban and Yang 1998), it is thought that the N-terminal domains must also interact for the complex to be fully functional (Ban and Yang 1998). Binding of ATP causes the proteins to undergo conformational changes, which are essential for the interaction between the N termini (Ban et al. 1999; Tran and Liskay 2000; Sacho et al. 2008). ATP hydrolysis and subsequent release of ADP is required to allow the protein complex to return to its initial state, completing the cycle so that the subunits are ready to bind ATP again if required. Using mutants of MLH1 and PMS1 that are presumed to be defective for ATP binding and/or ATP hydrolysis, it has been shown that both of these functions are essential for fully effective mismatch repair (Tran and Liskay 2000). However, the ATP binding and ATP hydrolysis mutants of PMS1 exhibited lower mitotic mutation rates than the corresponding MLH1 ATPase mutants, suggesting that there is functional asymmetry within the Mlh1p/Pms1p heterodimer (Tran and Liskay 2000; Hall et al. 2002). Another example of the asymmetry in the contributions of these subunits to function can be seen in assays that measure recombination between diverged sequences (homeologous recombination). The Mlh1p ATPase activity has been shown to be more important for the suppression of homeologous recombination than Pms1p ATPase activity (Welz-Voegele et al. 2002). This functional asymmetry is supported by in vitro biochemical analysis that demonstrated Pms1p has a lower ATP binding affinity than Mlh1p (Hall et al. 2002).As mentioned above, Mlh1p/Mlh3p function in the Msh4p/Msh5p pathway for meiotic recombination (Hunter and Borts 1997; Santucci-Darmanin et al. 2000). The Msh4p/Msh5p complex is thought to act in the stabilization of Holliday junction intermediates to allow their resolution in a crossover configuration (Snowden et al. 2004). The Mlh1p/Mlh3p complex has been suggested to act in the resolution of these structures, either directly or indirectly. Human Pms2 and its yeast homolog, Pms1p, have been shown to possess a latent endonuclease activity, conferred by a motif that is conserved among some of the MutL homologs, including Mlh3p (Kadyrov et al. 2006, 2007). Mutations in the DHQA(X)2E(X)4E motif in yeast MLH3 cause defects in both mismatch repair and meiotic recombination equivalent to mlh3Δ, suggesting that Mlh3p may also possess an endonuclease activity that is important for the generation of crossovers (Nishant et al. 2008).ATP binding by Mlh1p has been shown to be important for both of its meiotic functions (crossing over and repair of heteroduplex DNA) (Pang et al. 1997; Tran and Liskay 2000; Hoffmann et al. 2003). In contrast, the ATP hydrolysis mutant mlh1-E31A/mlh1-E31A appears to have no effect on meiotic recombination (Tran and Liskay 2000; Hoffmann et al. 2003). This may partly be explained by in vitro studies demonstrating that this mutant exhibits a low level of ATPase activity (Hall et al. 2002).The meiotic functions of MLH1 can be functionally separated as shown by mutating the same residue, G98, to different amino acids (Hoffmann et al. 2003). The residue G98 is situated in the ATPase motif in the GFRGEAL box (GYRGDAL in Mlh3p), which forms the lid of the ATP binding pocket. Mutations in this motif are predicted to affect ATP binding and/or heterodimerization with Pms1p (Ban and Yang 1998; Ban et al. 1999). Mutating the residue G98 in the ATP binding lid to alanine resulted in defective repair of heteroduplex DNA while crossing over was unaffected, but when the same residue was mutated to valine both mismatch repair and crossover functions were defective (Hoffmann et al. 2003). The mlh1-G98V mutant disrupts the interaction of Mlh1p with Pms1p, while mlh1-G98A does not (Pang et al. 1997). This may contribute to the difference observed in the effect on crossing over as Mlh1p is thought to interact with Pms1p and Mlh3p through the same residues (Wang et al. 1999; Kondo et al. 2001). Consequently if the interaction with Pms1p is affected then it is likely that the interaction with Mlh3p is also disrupted.We constructed mlh3 mutants corresponding to the ATP binding and ATP hydrolysis mutants of mlh1 to explore the role of Mlh3p in meiotic recombination. We also constructed mlh3-G97A and mlh3-G97V mutants, equivalent to the mlh1-G98A/V pair that has been shown to differentially affect the mitotic and meiotic functions of Mlh1p. All mutants were assayed for mitotic mismatch repair, meiotic heteroduplex repair, crossing over, and chromosome segregation.  相似文献   

8.
Studies of Mutations in T4 Control Genes 33 and 55   总被引:2,自引:0,他引:2       下载免费PDF全文
H. Robert Horvitz 《Genetics》1975,79(3):349-360
  相似文献   

9.
The essential Caenorhabditis elegans gene rfl-1 encodes one subunit of a heterodimeric E1-activating enzyme in the Nedd8 ubiquitin-like protein conjugation pathway. This pathway modifies the Cullin scaffolds of E3 ubiquitin ligases with a single Nedd8 moiety to promote ligase function. To identify genes that influence neddylation, we used a synthetic screen to identify genes that, when depleted with RNAi, enhance or suppress the embryonic lethality caused by or198ts, a temperature-sensitive (ts) mutation in rfl-1. We identified reproducible suppressor and enhancer genes and employed a systematic specificity analysis for each modifier using four unrelated ts embryonic lethal mutants. Results of this analysis highlight the importance of specificity controls in identifying genetic interactions relevant to a particular biological process because 8/14 enhancers and 7/21 suppressors modified lethality in other mutants. Depletion of the strongest specific suppressors rescued the early embryonic cell division defects in rfl-1(or198ts) mutants. RNAi knockdown of some specific suppressors partially restored Cullin neddylation in rfl-1(or198ts) mutants, consistent with their gene products normally opposing neddylation, and GFP fusions to several suppressors were detected in the cytoplasm or the nucleus, similar in pattern to Nedd8 conjugation pathway components in early embryonic cells. In contrast, depletion of the two strongest specific enhancers did not affect the early embryonic cell division defects observed in rfl-1(or198ts) mutants, suggesting that they may act at later times in other essential processes. Many of the specific modifiers are conserved in other organisms, and most are nonessential. Thus, when controlled properly for specificity, modifier screens using conditionally lethal C. elegans mutants can identify roles for nonessential but conserved genes in essential processes.UBIQUITIN-mediated proteolysis regulates many biological processes (Nandi et al. 2006). In the early Caenorhabditis elegans embryo, these include oocyte maturation, cell cycle progression, cell polarization, and cell fate patterning, all of which require the timely destruction of maternally expressed proteins (Bowerman and Kurz 2006; Greenstein and Lee 2006). One C. elegans protein targeted for proteolysis early in embryogenesis is MEI-1, the AAA-ATPase subunit of the microtubule-severing complex called katanin (Mains et al. 1990; Dow and Mains 1998; Srayko et al. 2000; Kurz et al. 2002; Pintard et al. 2003a; Xu et al. 2003). Katanin is a heterodimer of two subunits called p60 and p80 in vertebrates and MEI-1 and MEI-2 in C. elegans. Katanin in C. elegans is required for proper assembly and function of the small, barrel-shaped meiotic spindles (Albertson and Thomson 1993; McNally et al. 2006) and must be degraded after meiotic divisions to permit assembly of the much larger first mitotic spindle in the one-cell zygote. In mutants that fail to degrade katanin after the completion of meiosis, the first mitotic spindle is fragmented and mis-oriented, cytokinesis is defective, and the embryos die without hatching (Dow and Mains 1998; Srayko et al. 2000; Kurz et al. 2002).The katanin subunit MEI-1 is targeted for poly-ubiquitylation and proteolytic destruction by a Cullin-based E3 ligase (Kurz et al. 2002). This complex includes the Cullin scaffolding protein CUL-3 and a substrate-specific adaptor called MEL-26 that binds to CUL-3 through a BTB domain and to MEI-1 through a MATH domain (Pintard et al. 2003b). Cullin 3-based E3 ligases in mammals also utilize substrate-specific adaptor proteins that, like MEL-26, have both a Cullin-binding BTB/POZ domain and another protein–protein interaction domain that binds to the substrate (Geyer et al. 2003; Cullinan et al. 2004; Angers et al. 2006). While MEI-1/Katanin downregulation by the CUL-3/MEL-26 E3 ligase is essential at most growth temperatures, a mel-26 null mutation is viable at the low growth temperature of 15° (Lu and Mains 2007). This bypass of mel-26 at 15° depends at least in part on the anaphase-promoting complex and its targeting of MEI-1 for proteolytic degradation (Lu and Mains 2007). Phosphorylation by the kinase MBK-2 primes MEI-1 for proteolysis (Quintin et al. 2003; Stitzel et al. 2007) and also promotes the downregulation of MEI-1 by the anaphase-promoting complex (Lu and Mains 2007).CUL-3 is the only C. elegans Cullin thus far identified that requires modification by the ubiquitin-like protein Nedd8 (Bowerman and Kurz 2006). In contrast, C. elegans CUL-2 is required for progression through meiosis and for the localized degradation of cell fate determinants in one-cell-stage embryos (Liu et al. 2004; Sonneville and Gonczy 2004), but neddylation-defective mutants do not exhibit these early defects (Bowerman and Kurz 2006). Cullin neddylation is mediated by the Nedd8 protein conjugation pathway, which begins with a heterodimeric E1-activating enzyme consisting of ULA-1 and RFL-1 (Uba3p in budding yeast) and also includes the E2-conjugating enzyme UBC-12 (Jones and Candido 2000; Srayko et al. 2000; Kurz et al. 2002) and the E3 ligase DCN-1 (Kurz et al. 2005).The downregulation of MEI-1/katanin by the CUL-3/MEL-26 E3 ligase requires a balance of both CUL-3 neddylation, which is mediated by the Nedd8 conjugation pathway, and deneddylation, which is mediated by the conserved COP-9 Signalosome (Pintard et al. 2003a). Other Cullin-based E3 ubiquitin ligases also require a balance of neddylation and deneddylation (Lyapina et al. 2001; Schwechheimer et al. 2001; Bornstein et al. 2006; Hetfeld et al. 2008). Deneddylation may modulate activation of the E3 ligase and thereby prevent the premature degradation of substrate adaptor proteins that also can become poly-ubiquitylated and degraded as a result of E3 ligase function.To identify additional factors that influence neddylation, and the downregulation of MEI-1/katanin after the completion of meiosis in C. elegans, we report here our use of RNA interference (RNAi) to reduce gene functions in a temperature-sensitive (ts) neddylation-defective mutant, rfl-1(or198ts). The discovery of RNAi and its systemic properties in C. elegans have made it possible to systematically target C. elegans genes for depletion by feeding worms bacterial strains that express double-strand RNAs corresponding to C. elegans gene sequences (Fire et al. 1998; Timmons et al. 2001; Feinberg and Hunter 2003; Baugh et al. 2005; Lehner et al. 2006; van Haaften et al. 2006). Furthermore, chemical mutagenesis screens have identified temperature-sensitive mutations in many essential C. elegans genes, which can be used for synthetic screens by choosing intermediate-growth temperatures that sensitize the genetic background and also optimize visual scoring of embryonic viability. Recently, genomewide RNAi screens have been used to identify C. elegans genes that, when reduced in function, restore viability to temperature-sensitive, embryonic-lethal mutants (Labbe et al. 2006; O''Rourke et al. 2007). Because a loss of suppressor function restores mutant viability, the suppressors may negatively regulate either the wild-type gene product or the process that requires the wild-type gene product.Here we report our identification of C. elegans genes that, when reduced in function by feeding RNAi, reproducibly suppressed or enhanced rfl-1(or198ts) embryonic lethality. Most suppressors were specific for rfl-1(or198ts), while specific enhancement was less common. Many of the rfl-1-specific suppressors and enhancers are conserved but appear nonessential. GFP fusions to several specific suppressors exhibit localization patterns that resemble those known for neddylation pathway components, and depletion of some of these partially restored CUL-3 neddylation in rfl-1(or198ts) mutants. In addition to identifying possible roles for conserved genes in cullin neddylation, we report the first quantitative analysis of specificity for both the enhancement and the suppression of a conditionally lethal mutant in C. elegans. Our results highlight the importance of testing genetic modifiers of conditionally lethal mutants for locus specificity.  相似文献   

10.
11.
We previously demonstrated efficient l-valine production by metabolically engineered Corynebacterium glutamicum under oxygen deprivation. To achieve the high productivity, a NADH/NADPH cofactor imbalance during the synthesis of l-valine was overcome by engineering NAD-preferring mutant acetohydroxy acid isomeroreductase (AHAIR) and using NAD-specific leucine dehydrogenase from Lysinibacillus sphaericus. Lactate as a by-product was largely eliminated by disrupting the lactate dehydrogenase gene ldhA. Nonetheless, a few other by-products, particularly succinate, were still produced and acted to suppress the l-valine yield. Eliminating these by-products therefore was deemed key to improving the l-valine yield. By additionally disrupting the phosphoenolpyruvate carboxylase gene ppc, succinate production was effectively suppressed, but both glucose consumption and l-valine production dropped considerably due to the severely elevated intracellular NADH/NAD+ ratio. In contrast, this perturbed intracellular redox state was more than compensated for by deletion of three genes associated with NADH-producing acetate synthesis and overexpression of five glycolytic genes, including gapA, encoding NADH-inhibited glyceraldehyde-3-phosphate dehydrogenase. Inserting feedback-resistant mutant acetohydroxy acid synthase and NAD-preferring mutant AHAIR in the chromosome resulted in higher l-valine yield and productivity. Deleting the alanine transaminase gene avtA suppressed alanine production. The resultant strain produced 1,280 mM l-valine at a yield of 88% mol mol of glucose−1 after 24 h under oxygen deprivation, a vastly improved yield over our previous best.  相似文献   

12.
1. Human uterine cervical stroma was found to contain a Ca2+-independent neutral proteinase against casein and N-benzoyl-dl-arginine p-nitroanilide (Bz-dl-Arg-Nan). This enzyme was tightly bound to an insoluble material (20000g pellet) and was solubilized by high concentrations of NaCl or KCl. High concentrations of them in the reaction system, however, inhibited reversibly the activity of this enzyme. 2. The neutral proteinase was partially purified by extraction with NaCl, gel filtration on Sephadex G-200 and affinity chromatography on casein–Sepharose. 3. The optimal pH of this partially purified enzyme was 7.4–8.0 against casein and Bz-dl-Arg-Nan. The molecular weight of the enzyme was found to be about 1.4×105 by gel filtration on Sephadex G-200. 4. The enzyme was significantly inhibited by di-isopropyl phosphorofluoridate (0.1mm). High concentration of phenylmethanesulphonyl fluoride (5mm), 7-amino-1-chloro-3-l-tosylamidoheptan-2-one (0.5mm), antipain (10μm) or leupeptin (10μm) was also found to be inhibitory, but chymostatin (40μg/ml), soya-bean trypsin inhibitor (2.5mg/ml), human plasma (10%, v/v), p-chloromercuribenzoate (1mm), EDTA (10mm) and 1-chloro-4-phenyl-3-l-tosylamidobutan-2-one (1mm) had no effect on the enzyme. 5. The neutral proteinase hydrolysed casein, Bz-dl-Arg-Nan and heat-denatured collagen, but was inactive towards native collagen and several synthetic substrates, such as 4-phenylazobenzyloxycarbonyl-Pro-Leu-Gly-Pro-d-Arg, 3-carboxypropionyl-Ala-Ala-Ala p-nitroanilide and 2,4-dinitrophenyl-Pro-Gln-Gly-Ile-Ala-Gly-Gln-d-Arg, and also proteoglycan. The enzyme did not act as a plasminogen activator. 6. These properties suggested that a neutral proteinase in the human uterine cervix was different from enzymes previously reported.  相似文献   

13.
Early in the process of speciation, reproductive failures occur in hybrid animals between genetically diverged populations. The sterile hybrid animals are often males in mammals and they exhibit spermatogenic disruptions, resulting in decreased number and/or malformation of mature sperms. Despite the generality of this phenomenon, comparative study of phenotypes in hybrid males from various crosses has not been done, and therefore the comprehensive genetic basis of the disruption is still elusive. In this study, we characterized the spermatogenic phenotype especially during meiosis in four different cases of reproductive isolation: B6-ChrXMSM, PGN-ChrXMSM, (B6 × Mus musculus musculus-NJL/Ms) F1, and (B6 × Mus spretus) F1. The first two are consomic strains, both bearing the X chromosome of M. m. molossinus; in B6-ChrXMSM, the genetic background is the laboratory strain C57BL/6J (predominantly M. m. domesticus), while in PGN-ChrXMSM the background is the PGN2/Ms strain purely derived from wild M. m. domesticus. The last two cases are F1 hybrids between mouse subspecies or species. Each of the hybrid males exhibited cell-cycle arrest and/or apoptosis at either one or two of three distinct meiotic stages: premeiotic stage, zygotene-to-pachytene stage of prophase I, and metaphase I. This study shows that the sterility in hybrid males is caused by spermatogenic disruptions at multiple stages, suggesting that the responsible genes function in different cellular processes. Furthermore, the stages with disruptions are not correlated with the genetic distance between the respective parental strains.WHEN animals from genetically diverged populations hybridize, complete or partial sterility is often observed in the F1 hybrids or in their descendants. This phenomenon belonging to postzygotic reproductive isolation accelerates irreversible genetic divergence by preventing free gene flow across the two diverging populations, and thereby plays a pivotal role in speciation. Sexual dimorphism is a general feature of reproductive isolation (Wu and Davis 1993; Laurie 1997; Orr 1997; Kulathinal and Singh 2008). In mammals, impairment is much more severe in males than in females, and in general the heterogametic sex is more sensitive to interspecific and intersubspecific genetic incompatibility. This phenomenon is well known as Haldane''s rule (Haldane 1922; Laurie 1997; Orr 1997).In many animals, the reproductive isolation is caused by spermatogenic disruptions characterized by reduced number of germ cells and small testis size. These animals include Drosophila (Joly et al. 1997), stickleback fish Pungitius (Takahashi et al. 2005), caviomorph rodent Thrichomys (Borodin et al. 2006), house musk shrew Suncus (Borodin et al. 1998), wallaby Petrogale (Close et al. 1996), and genus Mus (Forejt and Iványi 1974; Matsuda et al. 1992; Hale et al. 1993; Yoshiki et al. 1993; Kaku et al. 1995; Gregorová and Forejt 2000; Elliott et al. 2001, 2004; Good et al. 2008). Although reproductive isolation by spermatogenic impairment is a well-known phenomenon, its underlying genetic mechanism and molecular basis have remained elusive. The Dobzhansky–Muller model, which infers that hybrid sterility or inviability is caused by deleterious epistatic interactions between nuclear genes derived from their respective parent species or subspecies (Dobzhansky 1936; Muller 1942), is widely accepted in animals and plants and is also applicable to the sterility of hybrid animals in F2 or backcross generations, so-called hybrid breakdown, in which the genes causing postzygotic reproductive isolation are partially recessive (Orr 2005).The genetic incompatibility between house mouse subspecies is an ideal animal model for studying the early stage of speciation. Two subspecies of mouse, Mus musculus domesticus and M. m. musculus, diverged from their common ancestor 0.3–1.0 MYA (Yonekawa et al. 1980; Moriwaki 1994; Bonhomme and Guénet 1996; Boursot et al. 1996; Din et al. 1996). M. m. domesticus ranges across western Europe and the Middle East, whereas M. m. musculus ranges throughout eastern Europe and northern Asia (Bonhomme and Guénet 1996). The two subspecies meet in a narrow hybrid zone, which is most likely maintained by a balance between dispersal and selection against hybrids (Hunt and Selander 1973; Bonhomme and Guénet 1996; Payseur et al. 2004). M. m. domesticus also displays reproductive isolation from the Japanese wild mouse, M. m. molossinus, which originated from hybridization of M. m. castaneus and M. m. musculus and its nuclear genome is predominantly derived from M. m. musculus (Yonekawa et al. 1980, 1988; Moriwaki 1994; Sakai et al. 2005). To investigate the reproductive isolation between M. m. domesticus and M. m. molossinus, we previously constructed a consomic strain B6-ChrXMSM (Oka et al. 2004). This strain has the X chromosome from the MSM/Ms strain, which is derived from M. m. molossinus, in the genetic background of the laboratory strain C57BL/6J (B6), which is predominantly derived from M. m. domesticus (Moriwaki 1994). F1 hybrid animals between B6 and MSM/Ms strains are fully fertile. On the contrary, B6-ChrXMSM shows male-specific sterility characterized by a reduced sperm number and dysfunction of the sperm, including abnormal morphology and low motility, indicating that B6-ChrXMSM is a model of hybrid breakdown in animals (Oka et al. 2004, 2007). Our previous study indicated that the abnormal morphology of the sperm head results from the genetic incompatibility between MSM/Ms-derived X-linked genes and B6 genes on autosomes including chromosomes 1 and 11 (Oka et al. 2007).In this study, to understand the genetic mechanism of reproductive isolation in mice, we first undertook in-depth characterization of phenotype for each B6-ChrXMSM male especially during meiosis. Meiosis is a special cell division that produces four haploid cells after one round of chromosome replication and two rounds of chromosome segregation. During meiosis, homologous chromosomes pair, synapse, undergo crossing over, and achieve bipolar attachment to the spindle to segregate one set of chromosomes to each daughter cell. Homologous recombination is initiated during the leptotene stage of meiotic prophase I with the formation of DNA double-strand breaks (DSBs), which are repaired immediately during the zygotene stage or after crossing over of homologous chromosomes during the pachytene stage (Roeder 1997; Tarsounas and Moens 2001).During the first wave of spermatogenesis, most mitotic spermatogonia in the B6-ChrXMSM testes fail to initiate meiotic DNA replication. Some proportion of those spermatogonia that enter into meiosis are again arrested and eliminated by apoptosis at the pachytene stage, resulting in the production of a small number of sperms. We extended the same analysis to three other cases of reproductive isolation. Another consomic strain PGN-ChrXMSM has an MSM/Ms-derived X chromosome in the genetic background of the PGN2/Ms strain derived from wild mice (M. m. domesticus). PGN-ChrXMSM males produce a small number of dysfunctional sperms as was the case with B6-ChrXMSM males, but the former males show apoptosis mainly at metaphase of meiosis I. Furthermore, we examined F1 hybrid males from intersubspecific cross of (B6 × M. m. musculus-NJL/Ms) and interspecific cross of (B6 × M. spretus). These F1 hybrid males exhibited apoptosis at metaphase I and at the zygotene-to-pachytene stage of prophase I. As a whole, the postzygotic reproductive isolation in mice is caused by disruptions at a minimum of three different spermatogenic stages.  相似文献   

14.
The first enzyme in the pathway for l-arabinose catabolism in eukaryotic microorganisms is a reductase, reducing l-arabinose to l-arabitol. The enzymes catalyzing this reduction are in general nonspecific and would also reduce d-xylose to xylitol, the first step in eukaryotic d-xylose catabolism. It is not clear whether microorganisms use different enzymes depending on the carbon source. Here we show that Aspergillus niger makes use of two different enzymes. We identified, cloned, and characterized an l-arabinose reductase, larA, that is different from the d-xylose reductase, xyrA. The larA is up-regulated on l-arabinose, while the xyrA is up-regulated on d-xylose. There is however an initial up-regulation of larA also on d-xylose but that fades away after about 4 h. The deletion of the larA gene in A. niger results in a slow growth phenotype on l-arabinose, whereas the growth on d-xylose is unaffected. The l-arabinose reductase can convert l-arabinose and d-xylose to their corresponding sugar alcohols but has a higher affinity for l-arabinose. The Km for l-arabinose is 54 ± 6 mm and for d-xylose 155 ± 15 mm.  相似文献   

15.
Evidence for a specific glutamate/h cotransport in isolated mesophyll cells   总被引:1,自引:1,他引:0  
Mechanically isolated Asparagus sprengeri Regel mesophyll cells were suspended in 1 millimolar CaSO4. Immediate alkalinization of the medium occured on the addition of 1 millimolar concentrations of l-glutamate (Glu) and its analog l-methionine-d,l-sulfoximine (l-MSO). d-Glu and the l isomers of the protein amino acids did not elicit alkalinization. l-Glu dependent alkalinization was transient and acidification resumed after approximately 30 to 45 minutes. At pH 6.0, 5 millimolar l-Glu stimulated initial rates of alkalinization that varied between 1.3 to 4.1 nmol H+/106 cells·minute. l-Glu dependent alkalinization was saturable, increased with decreasing pH, was inhibited by carbonyl cyanide-p-trichloromethoxyphenyl hydrazone (CCCP), and was not stimulated by light. Uptake of l-[U-14C]glutamate increased as the pH decreased from 6.5 to 5.5, and was inhibited by l-MSO. l-Glu had no influence on K+ efflux. Although evidence for multiple amino acid/proton cotransport systems has been found in other tissues, the present report indicates that a highly specific l-Glu/proton uptake process is present in Asparagus mesophyll cells.  相似文献   

16.
While mitochondria are renowned for their role in energy production, they also perform several other integral functions within the cell. Thus, it is not surprising that mitochondrial dysfunction can negatively impact cell viability. Although mitochondria have received an increasing amount of attention in recent years, there is still relatively little information about how proper maintenance of mitochondria and its genomes is achieved. The Neurospora crassa mus-10 mutant was first identified through its increased sensitivity to methyl methanesulfonate (MMS) and was thus believed to be defective in some aspect of DNA repair. Here, we report that mus-10 harbors fragmented mitochondria and that it accumulates deletions in its mitochondrial DNA (mtDNA), suggesting that the mus-10 gene product is involved in mitochondrial maintenance. Interestingly, mus-10 begins to senesce shortly after deletions are visualized in its mtDNA. To uncover the function of MUS-10, we used a gene rescue approach to clone the mus-10 gene and discovered that it encodes a novel F-box protein. We show that MUS-10 interacts with a core component of the Skp, Cullin, F-box containing (SCF) complex, SCON-3, and that its F-box domain is essential for its function in vivo. Thus, we provide evidence that MUS-10 is part of an E3 ubiquitin ligase complex involved in maintaining the integrity of mitochondria and may function to prevent cellular senescence.THE mus-10 mutant was isolated from a screen aimed at identifying Neurospora crassa strains that were sensitive to MMS and therefore likely to lack proper DNA repair mechanisms (Kafer and Perlmutter 1980). Epistasis analyses involving mus-10 suggested that it belonged to the uvs-6 epistasis group, which functions in recombination repair (Kafer and Perlmutter 1980; Kafer 1983). However, mus-10 did not display several phenotypes common to other members of the uvs-6 epistasis group: chromosomal instability, a high sensitivity to histidine, and the inability to produce viable ascospores in homozygous crosses (Newmeyer et al. 1978; Newmeyer and Galeazzi 1978; Kafer and Perlmutter 1980; Kafer 1981; Schroeder 1986; Watanabe et al. 1997; Handa et al. 2000; Sakuraba et al. 2000). Furthermore, the frequencies of spontaneous and radiation-induced mutation observed in mus-10 were similar to those of a wild-type strain (Kafer 1981). Past efforts to uncover the nature of these discrepancies or the function of the mus-10 gene product have been uninformative.The majority of cellular ATP is produced in mitochondria through aerobic respiration, which couples electron flow through respiratory complexes within the mitochondrial inner membrane with oxidative phosphorylation. Besides their role in ATP synthesis, mitochondria are also involved in many other cellular processes including beta-oxidation (Bartlett and Eaton 2004), calcium homeostasis (Gunter et al. 2004; Rimessi et al. 2008), production of iron-sulfur clusters (Zheng et al. 1998; Gerber and Lill 2002; Lill and Muhlenhoff 2005; Rouault and Tong 2005), and apoptosis (Green 2005; Antignani and Youle 2006; Xu and Shi 2007). Although virtually all mitochondrial proteins are encoded within the nucleus, a small number of proteins are encoded by mitochondrial DNA (mtDNA). The integrity of the mitochondrial genome may affect cell survival as mutations in mtDNA accumulate in patients suffering from severe neurological diseases including Alzheimer''s, Huntington''s and Parkinson''s, as well as several types of cancer (Chatterjee et al. 2006; Higuchi 2007; Krishnan et al. 2007; Reeve et al. 2008). The number of mtDNA mutations also increases with age, suggesting a link between mitochondrial dysfunction and ageing (Cortopassi and Arnheim 1990; Corral-Debrinski et al. 1992; Cortopassi et al. 1992; Simonetti et al. 1992; Reeve et al. 2008). Contrary to the single genome in the nucleus, there are several copies of mtDNA in each mitochondrion. Thus, defects in a few mitochondrial genomes do not necessarily lead to mitochondrial dysfunction. Many patients suffering from mitochondrial diseases exhibit heteroplasmy, a phenomenon in which a mixture of wild-type and mutant mtDNAs exist in a single cell. The ratio of wild-type to mutant mtDNAs is critical in determining the penetrance of the genetic defect, where mutant loads >60% are required to cause respiratory chain dysfunction within an individual cell (Boulet et al. 1992; Chomyn et al. 1992; Sciacco et al. 1994).Even though N. crassa strains are generally deemed immortal if they can be subcultured ∼50 times, a wild-type strain was recently reported to senesce after 12,000 hr of growth, implying that this fungus undergoes natural or programmed ageing (Maheshwari and Navaraj 2008; Kothe et al. 2010). However, replicative life span is also influenced by genetic background as certain mutations can cause progressive deterioration of growth, ultimately leading to death. One such example is the nuclear-encoded natural death (nd), which when mutant causes a senescence phenotype correlating with the accumulation of multiple mtDNA deletions (Sheng 1951; Seidel-Rogol et al. 1989). The deletions of mtDNA in nd occurred between two 70- to 701-bp direct repeats, suggesting that the nd gene product regulates recombination, repair, or replication of mtDNA (Bertrand et al. 1993). Another nuclear mutation, senescence (sen), was isolated from N. intermedia and introgressed into N. crassa (Navaraj et al. 2000). Deletions were also observed in the mtDNA of sen mutants, but unlike those occurring in nd were flanked by 6- to 10-bp repeats typically associated with GC-rich palindromic sequences (D''Souza et al. 2005). The nature of the sequences that flanked the mtDNA deletions in these two mutants supported the existence of two distinct systems of mtDNA recombination in N. crassa: a general system of homologous recombination (system I) and a site-specific mechanism (system II), mediated in part by nd and sen, respectively (Bertrand et al. 1993; D''Souza et al. 2005). The nd and sen mutations have been mapped to linkage groups I and V, respectively, but neither gene has been cloned and the precise function of their gene products remains unclear. Two ultraviolet (UV)-sensitive mutants, uvs-4 and uvs-5, are thought to undergo senescence, but unfortunately, these strains have not been studied in great detail (Schroeder 1970; Perkins et al. 1993; Hausner et al. 2006). Premature senescence has also been observed in cytoplasmic mutants of N. crassa including the E35 and ER-3 stopper mutants that harbor large mtDNA deletions, as well as strains that accumulate mitochondrial plasmids capable of inserting into mtDNA through homologous recombination (de Vries et al. 1986; Akins et al. 1989; Myers et al. 1989; Niagro and Mishra 1989; Court et al. 1991; Alves and Videira 1998).While trying to establish the role of MUS-10 in DNA repair, we discovered that the mus-10 mutant exhibited a shortened life span, an abnormal mitochondrial morphology and mtDNA instability. We cloned the mus-10 gene through its ability to complement the MMS sensitivity of the mus-10 mutant and revealed that it encoded a novel F-box protein. This suggested that MUS-10 is part of an Skp, Cullin, F-box containing (SCF) E3 ubiquitin ligase complex that targets proteins for degradation by the 26S proteasome. The data we present in this article offer proof that an SCF complex can regulate both mitochondrial maintenance and cellular senescence.  相似文献   

17.
We present the results of surveys of diversity in sets of >40 X-linked and autosomal loci in samples from natural populations of Drosophila miranda and D. pseudoobscura, together with their sequence divergence from D. affinis. Mean silent site diversity in D. miranda is approximately one-quarter of that in D. pseudoobscura; mean X-linked silent diversity is about three-quarters of that for the autosomes in both species. Estimates of the distribution of selection coefficients against heterozygous, deleterious nonsynonymous mutations from two different methods suggest a wide distribution, with coefficients of variation greater than one, and with the average segregating amino acid mutation being subject to only very weak selection. Only a small fraction of new amino acid mutations behave as effectively neutral, however. A large fraction of amino acid differences between D. pseudoobscura and D. affinis appear to have been fixed by positive natural selection, using three different methods of estimation; estimates between D. miranda and D. affinis are more equivocal. Sources of bias in the estimates, especially those arising from selection on synonymous mutations and from the choice of genes, are discussed and corrections for these applied. Overall, the results show that both purifying selection and positive selection on nonsynonymous mutations are pervasive.SURVEYS of DNA sequence diversity and divergence are shedding light on a number of questions in evolutionary genetics (for recent reviews, see Akey 2009; Sella et al. 2009). Two of the most important questions of this kind concern the distribution of selection coefficients against deleterious mutations affecting protein sequences and the proportion of amino acid sequence differences between related species that have been fixed by positive selection. Several different methods have been proposed for studying each of these questions, using different features of data on polymorphism and divergence at nonsynonymous and silent sites.For example, the parameters of the distribution of selection coefficients against deleterious amino acid mutations have been estimated by contrasting the numbers of nonsynonymous and silent within-species polymorphisms and fixed differences between species (Sawyer and Hartl 1992; Bustamante et al. 2002; Piganeau and Eyre-Walker 2003; Sawyer et al. 2007); by fitting the frequency spectra of nonsynonymous and silent variants to models of selection, mutation, and drift (Akashi 1999; Eyre-Walker et al. 2006; Keightley and Eyre-Walker 2007; Kryukov et al. 2007; Boyko et al. 2008; Eyre-Walker and Keightley 2009); or by comparing levels of nonsynonymous and silent diversities between species with different population sizes (Loewe and Charlesworth 2006; Loewe et al. 2006). The results of these different approaches generally agree in suggesting that there is a wide distribution of selection coefficients against nonsynonymous mutations and that the mean selection coefficient against heterozygous carriers of such mutations is very small. The results imply that a typical individual from a human population carries several hundred weakly deleterious mutations (Eyre-Walker et al. 2006; Kryukov et al. 2007; Boyko et al. 2008); for a typical Drosophila population, with its much higher level of variability, the number is probably an order of magnitude greater (Loewe et al. 2006; Keightley and Eyre-Walker 2007).The presence of this large load of slightly deleterious mutations in human and natural populations, most of which are held at low frequencies by natural selection, has many implications. From the point of view of understanding human genetic disease, it means that we have to face the likelihood that susceptibility to a disease can be influenced by variants at many loci, each with small effects (Kryukov et al. 2007). The pervasive presence of deleterious mutations throughout the genome contributes to inbreeding depression (Charlesworth and Willis 2009) and may mean that the effective population size is reduced by background selection effects, even in regions of the genome with normal levels of genetic recombination (Loewe and Charlesworth 2007). Their presence may contribute so strongly to Hill–Robertson effects (Hill and Robertson 1966; Felsenstein 1974) that they cause severely reduced levels of diversity and adaptation in low-recombination regions of the genome (Charlesworth et al. 2010) and create a selective advantage to maintaining nonzero levels of recombination (Keightley and Otto 2006; Charlesworth et al. 2010). In addition, having an estimate of the distribution of selection coefficients against deleterious nonsynonymous mutations allows their contribution to between-species divergence to be predicted, providing a way of estimating the fraction of fixed nonsynonymous differences caused by positive selection (Loewe et al. 2006; Boyko et al. 2008; Eyre-Walker and Keightley 2009).It is thus important to collect data that shed light on the properties of selection against nonsynonymous mutations in a wide range of systems and also to compare the results from different methods of estimation, since they are subject to different sources of difficulty and biases. In a previous study, we proposed the use of a comparison between two related species with different effective population sizes for this purpose (Loewe and Charlesworth 2006; Loewe et al. 2006), using Drosophila miranda and D. pseudoobscura as material. These are well suited for this type of study, as they are closely related, live together in similar habitats, and yet have very different levels of silent nucleotide diversity, indicating different effective population sizes (Ne). This study was hampered by our inability to compare the same set of loci across the two species and by the small number of loci that could be used. We here present the results of a much larger study of DNA variation at X-linked and autosomal loci for these two species, using D. affinis as a basis for estimating divergence. We compare the results, applying the method of Loewe et al. (2006) with that of Eyre-Walker and Keightley (2009) for estimating the distribution of deleterious selection coefficients and with McDonald–Kreitman test-based methods for estimating the proportion of nonsynonymous differences fixed by positive selection. While broadly confirming the conclusions from earlier studies, we note some possible sources of bias and describe methods for minimizing their effects.  相似文献   

18.
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   

19.
A Mg2+-dependent, alkaline phosphatase has been isolated from mature pollen of Lilium longiflorum Thunb., cv. Ace and partially purified. It hydrolyzes 1l- and 1d-myo-inositol 1-phosphate, myo-inositol 2-phosphate, and β-glycerophosphate at rates decreasing in the order named. The affinity of the enzyme for 1l- and 1d-myo-inositol 1-phosphate is approximately 10-fold greater than its affinity for myo-inositol 2-phosphate. Little or no activity is found with phytate, d-glucose 6-phosphate, d-glucose 1-phosphate, d-fructose 1-phosphate, d-fructose 6-phosphate, d-mannose 6-phosphate, or p-nitrophenyl phosphate. 3-Phosphosphoglycerate is a weak competitive inhibitor. myo-Inositol does not inhibit the reaction. Optimal activity is obtained at pH 8.5 and requires the presence of Mg2+. At 4 millimolar, Co2+, Fe2+ or Mn2+ are less effective. Substantial inhibition is obtained with 0.25 molar Li+. With β-glycerophosphate as substrate the Km is 0.06 millimolar and the reaction remains linear at least 2 hours. In 0.1 molar Tris, β-glycerophosphate yields equivalent amounts of glycerol and inorganic phosphate, evidence that transphosphorylation does not occur.  相似文献   

20.
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   

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