首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
《Autophagy》2013,9(11):1557-1576
L-arginine (L-Arg) deficiency results in decreased T-cell proliferation and impaired T-cell function. Here we have found that L-Arg depletion inhibited expression of different membrane antigens, including CD247 (CD3ζ), and led to an ER stress response, as well as cell cycle arrest at G0/G1 in both human Jurkat and peripheral blood mitogen-activated T cells, without undergoing apoptosis. By genetic and biochemical approaches, we found that L-Arg depletion also induced autophagy. Deprivation of L-Arg induced EIF2S1 (eIF2α), MAPK8 (JNK), BCL2 (Bcl-2) phosphorylation, and displacement of BECN1 (Beclin 1) binding to BCL2, leading to autophagosome formation. Silencing of ERN1 (IRE1α) prevented the induction of autophagy as well as MAPK8 activation, BCL2 phosphorylation and XBP1 splicing, whereas led T lymphocytes to apoptosis under L-Arg starvation, suggesting that the ERN1-MAPK8 pathway plays a major role in the activation of autophagy following L-Arg depletion. Autophagy was required for survival of T lymphocytes in the absence of L-Arg, and resulted in a reversible process. Replenishment of L-Arg made T lymphocytes to regain the normal cell cycle profile and proliferate, whereas autophagy was inhibited. Inhibition of autophagy by ERN1, BECN1 and ATG7 silencing, or by pharmacological inhibitors, promoted cell death of T lymphocytes incubated in the absence of L-Arg. Our data indicate for the first time that depletion of L-Arg in T lymphocytes leads to a reversible response that preserves T lymphocytes through ER stress and autophagy, while remaining arrested at G0/G1. Our data also show that the L-Arg depletion-induced ER stress response could lead to apoptosis when autophagy is blocked.  相似文献   

2.
Intracellular thiols like L-cystine and L-cystine play a critical role in the regulation of cellular processes. Here we show that Escherichia coli has two L-cystine transporters, the symporter YdjN and the ATP-binding cassette importer FliY-YecSC. These proteins import L-cystine, an oxidized product of L-cystine from the periplasm to the cytoplasm. The symporter YdjN, which is expected to be a new member of the L-cystine regulon, is a low affinity L-cystine transporter (K m = 1.1 μM) that is mainly involved in L-cystine uptake from outside as a nutrient. E. coli has only two L-cystine importers because ΔydjNΔyecS mutant cells are not capable of growing in the minimal medium containing L-cystine as a sole sulfur source. Another protein YecSC is the FliY-dependent L-cystine transporter that functions cooperatively with the L-cystine transporter YdeD, which exports L-cystine as reducing equivalents from the cytoplasm to the periplasm, to prevent E. coli cells from oxidative stress. The exported L-cystine can reduce the periplasmic hydrogen peroxide to water, and then generated L-cystine is imported back into the cytoplasm via the ATP-binding cassette transporter YecSC with a high affinity to L-cystine (K m = 110 nM) in a manner dependent on FliY, the periplasmic L-cystine-binding protein. The double disruption of ydeD and fliY increased cellular levels of lipid peroxides. From these findings, we propose that the hydrogen peroxide-inducible L-cystine/L-cystine shuttle system plays a role of detoxification of hydrogen peroxide before lipid peroxidation occurs, and then might specific prevent damage to membrane lipids.  相似文献   

3.
The tremendous social and economic impact of thrombotic disorders, together with the considerable risks associated to the currently available therapies, prompt for the development of more efficient and safer anticoagulants. Novel peptide-based thrombin inhibitors were identified using in silico structure-based design and further validated in vitro. The best candidate compounds contained both l- and d-amino acids, with the general sequence d-Phe(P3)-Pro(P2)-d-Arg(P1)-P1′-CONH2. The P1′ position was scanned with l- and d-isomers of natural or unnatural amino acids, covering the major chemical classes. The most potent non-covalent and proteolysis-resistant inhibitors contain small hydrophobic or polar amino acids (Gly, Ala, Ser, Cys, Thr) at the P1′ position. The lead tetrapeptide, d-Phe-Pro-d-Arg-d-Thr-CONH2, competitively inhibits α-thrombin''s cleavage of the S2238 chromogenic substrate with a Ki of 0.92 µM. In order to understand the molecular details of their inhibitory action, the three-dimensional structure of three peptides (with P1′ l-isoleucine (fPrI), l-cysteine (fPrC) or d-threonine (fPrt)) in complex with human α-thrombin were determined by X-ray crystallography. All the inhibitors bind in a substrate-like orientation to the active site of the enzyme. The contacts established between the d-Arg residue in position P1 and thrombin are similar to those observed for the l-isomer in other substrates and inhibitors. However, fPrC and fPrt disrupt the active site His57-Ser195 hydrogen bond, while the combination of a P1 d-Arg and a bulkier P1′ residue in fPrI induce an unfavorable geometry for the nucleophilic attack of the scissile bond by the catalytic serine. The experimental models explain the observed relative potency of the inhibitors, as well as their stability to proteolysis. Moreover, the newly identified direct thrombin inhibitors provide a novel pharmacophore platform for developing antithrombotic agents by exploring the conformational constrains imposed by the d-stereochemistry of the residues at positions P1 and P1′.  相似文献   

4.
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   

5.
Gd1a, Gd1b and Gt1b gangliosides were dispersed in the following membrane-mimicking systems: (a) homogeneous micelles; (b) mixed micelles with Gm1 ganglioside (which is resistant to the enzyme action), Triton X-100 or bovine serum albumin; (c) small unilamellar vesicles of egg phosphatidylcholine. The effect of dispersion on sialic acid release by Vibrio cholerae sialidase was studied. As reference substrates freely interacting with the enzyme the lipid-free carbohydrates of Gd1a and 3′-sialosyl-lactose were employed. The apparent Vmax. of the enzyme was, with all the gangliosides, dependent on the type of ganglioside dispersion. It was lowest for homogeneous micelles and mixed micelles with ganglioside Gm1, and increased about 6-fold for ganglioside/bovine serum albumin lipoprotein micelles, 15-fold for mixed-ganglioside/Triton X-100 micelles (optimal molar ratio 1:7.5) and 30-fold for phosphatidylcholine vesicles containing 2.5 mol% ganglioside (this proportion was optimal for enzyme activity on the vesicles). For ganglioside Gd1a, the activity on Triton X-100 mixed micelles and on mixed vesicles was even greater (3- and 6-fold respectively) than that displayed on Gd1a lipid-free carbohydrate. With each of the used gangliosides the apparent Km values were very similar values for homogeneous micelles and vesicular dispersions, but showed marked increases for Triton X-100 mixed micelles, approaching the values exhibited by reference oligosaccharides. Triton X-100 micelles and phosphatidylcholine vesicles did not appreciably alter the kinetics of sialidase action on 3′-sialosyl-lactose and on Gd1a lipid-free carbohydrate, indicating that the above effects are dependent on the intrinsic characteristics of the membrane-like systems containing gangliosides.  相似文献   

6.
7.
Autophagy is an evolutionarily conserved degradative pathway that has been implicated in a number of physiological events important for human health. This process was originally identified as a response to nutrient deprivation and is thought to serve in a recycling capacity during periods of nutritional stress. Autophagy activity appears to be highly regulated and multiple signaling pathways are known to target a complex of proteins that contains the Atg1 protein kinase. The data here extend these observations and identify a particular phosphorylation event on Atg1 as a potential control point within the autophagy pathway in Saccharomyces cerevisiae. This phosphorylation occurs at a threonine residue, T226, within the Atg1 activation loop that is conserved in all Atg1 orthologs. Replacing this threonine with a nonphosphorylatable residue resulted in a loss of Atg1 protein kinase activity and a failure to induce autophagy. This phosphorylation required the presence of a functional Atg1 kinase domain and two known regulators of Atg1 activity, Atg13 and Atg17. Interestingly, the levels of this modification were found to increase dramatically upon exposure to conditions that induce autophagy. In addition, T226 phosphorylation was associated with an autophosphorylated form of Atg1 that was found specifically in cells undergoing the autophagy process. In all, these data suggest that autophosphorylation within the Atg1 activation loop may represent a point of regulatory control for this degradative process.MACROAUTOPHAGY (hereafter referred to as autophagy) is a highly conserved process of self-degradation that is essential for cell survival during periods of nutrient limitation (Tsukada and Ohsumi 1993). During autophagy, a double membrane grows out from a specific nucleation site, known as the pre-autophagosomal structure, or PAS, in Saccharomyces cerevisiae and the phagophore assembly site in mammals (Suzuki and Ohsumi 2007). This membrane encapsulates bulk protein and other constituents of the cytoplasm and ultimately targets this material to the vacuole/lysosome for degradation (Xie and Klionsky 2007). Recent studies have linked this pathway to a number of processes important for human health, including tumor suppression, innate immunity, and neurological disorders, like Huntington''s disease (Rubinsztein et al. 2007; Levine and Kroemer 2008). Determining how this pathway is regulated is therefore important for our understanding of these processes and our attempts to manipulate autophagy in clinically beneficial ways.Most of the molecular components of the autophagy pathway were initially characterized in the budding yeast, S. cerevisiae, but orthologs of many of these Atg proteins have since been found in other eukaryotes (Tsukada and Ohsumi 1993; Meijer et al. 2007). A complex of proteins that contains the Atg1 protein kinase is of special interest and appears to be a key point of regulatory control within this pathway (Kamada et al. 2000; Budovskaya et al. 2005; He and Klionsky 2009; Stephan et al. 2009). In S. cerevisiae, genetic and biochemical data indicate that this complex is targeted by at least three different signaling pathways. Two of these pathways, involving the Tor and cAMP-dependent protein kinases, inhibit this process, whereas the AMP-activated protein kinase is needed for the full induction of autophagy (Noda and Ohsumi 1998; Wang et al. 2001; Budovskaya et al. 2004; Stephan and Herman 2006; Kamada et al. 2010). The manner in which these signaling pathways regulate Atg1 activity and the precise role of this kinase in the autophagy process are presently matters of intense scrutiny.Although Atg1 kinase activity is required for the induction of autophagy, relatively little is known about how this enzyme is regulated in vivo. Two proteins associated with Atg1, Atg13 and Atg17, have been shown to be required for full Atg1 kinase activity both in vitro and in vivo (Kamada et al. 2000; Stephan et al. 2009). The roles of these proteins appear to be conserved through evolution as functional homologs of both have been identified in fruit flies and/or mammals (Hara et al. 2008; Chan et al. 2009; Chang and Neufeld 2009; Ganley et al. 2009; Hosokawa et al. 2009; Jung et al. 2009; Mercer et al. 2009). However, it is not yet clear precisely how these proteins stimulate Atg1 activity. In this study, we show that Atg1 is autophosphorylated within the activation loop and that this phosphorylation is required for both Atg1 kinase activity and the induction of autophagy. The activation loop is a structurally conserved element within the kinase domain and phosphorylation within this loop is often a necessary prerequisite for efficient substrate binding and/or phosphotransfer in the catalytic site (Johnson et al. 1996; Nolen et al. 2004). This loop generally corresponds to the sequence between two signature elements within the core kinase domain, the DFG and APE motifs (Hanks and Hunter 1995). Phosphorylation within this loop tends to result in a more ordered structure for this region and the proper positioning of key elements within the catalytic core of the kinase domain (Knighton et al. 1991; Johnson and O''reilly 1996; Huse and Kuriyan 2002). We found that Atg1 activation loop phosphorylation was correlated with the onset of autophagy and that replacing the site of phosphorylation with a phosphomimetic residue led to constitutive Atg1 autophosphorylation in vivo. In all, the data here suggest that Atg1 phosphorylation within its activation loop may be an important point of regulation within the autophagy pathway and models that discuss these data are presented.  相似文献   

8.
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   

9.
10.
11.
12.
13.
Joshua S. Paul  Yun S. Song 《Genetics》2010,186(1):321-338
The multilocus conditional sampling distribution (CSD) describes the probability that an additionally sampled DNA sequence is of a certain type, given that a collection of sequences has already been observed. The CSD has a wide range of applications in both computational biology and population genomics analysis, including phasing genotype data into haplotype data, imputing missing data, estimating recombination rates, inferring local ancestry in admixed populations, and importance sampling of coalescent genealogies. Unfortunately, the true CSD under the coalescent with recombination is not known, so approximations, formulated as hidden Markov models, have been proposed in the past. These approximations have led to a number of useful statistical tools, but it is important to recognize that they were not derived from, though were certainly motivated by, principles underlying the coalescent process. The goal of this article is to develop a principled approach to derive improved CSDs directly from the underlying population genetics model. Our approach is based on the diffusion process approximation and the resulting mathematical expressions admit intuitive genealogical interpretations, which we utilize to introduce further approximations and make our method scalable in the number of loci. The general algorithm presented here applies to an arbitrary number of loci and an arbitrary finite-alleles recurrent mutation model. Empirical results are provided to demonstrate that our new CSDs are in general substantially more accurate than previously proposed approximations.THE probability of observing a sample of DNA sequences under a given population genetics model—which is referred to as the sampling probability or likelihood—plays an important role in a wide range of problems in a genetic variation study. When recombination is involved, however, obtaining an analytic formula for the sampling probability has hitherto remained a challenging open problem (see Jenkins and Song 2009, 2010 for recent progress on this problem). As such, much research (Griffiths and Marjoram 1996; Kuhner et al. 2000; Nielsen 2000; Stephens and Donnelly 2000; Fearnhead and Donnelly 2001; De Iorio and Griffiths 2004a,b; Fearnhead and Smith 2005; Griffiths et al. 2008; Wang and Rannala 2008) has focused on developing Monte Carlo methods on the basis of the coalescent with recombination (Griffiths 1981; Kingman 1982a,b; Hudson 1983), a well-established mathematical framework that models the genealogical history of sample chromosomes. These Monte Carlo-based full-likelihood methods mark an important development in population genetics analysis, but a well-known obstacle to their utility is that they tend to be computationally intensive. For a whole-genome variation study, approximations are often unavoidable, and it is therefore important to think of ways to minimize the trade-off between scalability and accuracy.A popular likelihood-based approximation method that has had a significant impact on population genetics analysis is the following approach introduced by Li and Stephens (2003): Given a set Φ of model parameters (e.g., mutation rate, recombination rate, etc.), the joint probability p(h1, … , hn | Φ) of observing a set {h1, … , hn} of haplotypes sampled from a population can be decomposed as a product of conditional sampling distributions (CSDs), denoted by π,(1)where π(hk+1|h1, …, hk, Φ) is the probability of an additionally sampled haplotype being of type hk+1, given a set of already observed haplotypes h1, …, hk. In the presence of recombination, the true CSD π is unknown, so Li and Stephens proposed using an approximate CSD in place of π, thus obtaining the following approximation of the joint probability:(2)Li and Stephens referred to this approximation as the product of approximate conditionals (PAC) model. In general, the closer is to the true CSD π, the more accurate the PAC model becomes. Notable applications and extensions of this framework include estimating crossover rates (Li and Stephens 2003; Crawford et al. 2004) and gene conversion parameters (Gay et al. 2007; Yin et al. 2009), phasing genotype data into haplotype data (Stephens and Scheet 2005; Scheet and Stephens 2006), imputing missing data to improve power in association mapping (Stephens and Scheet 2005; Li and Abecasis 2006; Marchini et al. 2007; Howie et al. 2009), inferring local ancestry in admixed populations (Price et al. 2009), inferring human colonization history (Hellenthal et al. 2008), inferring demography (Davison et al. 2009), and so on.Another problem in which the CSD plays a fundamental role is importance sampling of genealogies under the coalescent process (Stephens and Donnelly 2000; Fearnhead and Donnelly 2001; De Iorio and Griffiths 2004a,b; Fearnhead and Smith 2005; Griffiths et al. 2008). In this context, the optimal proposal distribution can be written in terms of the CSD π (Stephens and Donnelly 2000), and as in the PAC model, an approximate CSD may be used in place of π. The performance of an importance sampling scheme depends critically on the proposal distribution and therefore on the accuracy of the approximation . Often in conjunction with composite-likelihood frameworks (Hudson 2001; Fearnhead and Donnelly 2002), importance sampling has been used in estimating fine-scale recombination rates (McVean et al. 2004; Fearnhead and Smith 2005; Johnson and Slatkin 2009).So far, a significant scope of intuition has gone into choosing the approximate CSDs used in these problems (Marjoram and Tavaré 2006). In the case of completely linked loci, Stephens and Donnelly (2000) suggested constructing an approximation by assuming that the additional haplotype hk+1 is an imperfect copy of one of the first k haplotypes, with copying errors corresponding to mutation. Fearnhead and Donnelly (2001) generalized this construction to include crossover recombination, assuming that the haplotype hk+1 is an imperfect mosaic of the first k haplotypes (i.e., hk+1 is obtained by copying segments from h1, …, hk, where crossover recombination can change the haplotype from which copying is performed). The associated CSD, which we denote by , can be interpreted as a hidden Markov model and so admits an efficient dynamic programming solution. Finally, Li and Stephens (2003) proposed a modification to Fearnhead and Donnelly''s model that limits the hidden state space, thereby providing a computational simplification; we denote the corresponding approximate CSD by .Although these approaches are computationally appealing, it is important to note that they are not derived from, though are certainly motivated by, principles underlying typical population genetics models, in particular the coalescent process (Griffiths 1981; Kingman 1982a,b; Hudson 1983). The main objective of this article is to develop a principled technique to derive an improved CSD directly from the underlying population genetics model. Rather than relying on intuition, we base our work on mathematical foundation. The theoretical framework we employ is the diffusion process. De Iorio and Griffiths (2004a,b) first introduced the diffusion-generator approximation technique to obtain an approximate CSD in the case of a single locus (i.e., no recombination). Griffiths et al. (2008) later extended the approach to two loci to include crossover recombination, assuming a parent-independent mutation model at each locus. In this article, we extend the framework to develop a general algorithm that applies to an arbitrary number of loci and an arbitrary finite-alleles recurrent mutation model.Our work can be summarized as follows. Using the diffusion-generator approximation technique, we derive a recursion relation satisfied by an approximate CSD. This recursion can be used to construct a closed system of coupled linear equations, in which the conditional sampling probability of interest appears as one of the unknown variables. The system of equations can be solved using standard numerical analysis techniques. However, the size of the system grows superexponentially with the number of loci and, consequently, so does the running time. To remedy this drawback, we introduce additional approximations to make our approach scalable in the number of loci. Specifically, the recursion admits an intuitive genealogical interpretation, and, on the basis of this interpretation, we propose modifications to the recursion, which then can be easily solved using dynamic programming. The computational complexity of the modified algorithm is polynomial in the number of loci, and, importantly, the resulting CSD has little loss of accuracy compared to that following from the full recursion.The accuracy of approximate CSDs has not been discussed much in the literature, except in the application-specific context for which they are being employed. In this article, we carry out an empirical study to explicitly test the accuracy of various CSDs and demonstrate that our new CSDs are in general substantially more accurate than previously proposed approximations. We also consider the PAC framework and show that our approximations also produce more accurate PAC-likelihood estimates. We note that for the maximum-likelihood estimation of recombination rates, the actual value of the likelihood may not be so important, as long as it is maximized near the true recombination rate. However, in many other applications—e.g., phasing genotype data into haplotype data, imputing missing data, importance sampling, and so on—the accuracy of the CSD and PAC-likelihood function over a wide range of parameter values may be important. Thus, we believe that the theoretical work presented here will have several practical implications; our method can be applied in a wide range of statistical tools that use CSDs, improving their accuracy.The remainder of this article is organized as follows. To provide intuition for the ensuing mathematics, we first describe a genealogical process that gives rise to our CSD. Using our genealogical interpretation, we consider two additional approximations and relate these to previously proposed CSDs. Then, in the following section, we derive our CSD using the diffusion-generator approach and provide mathematical statements for the additional approximations; some interesting limiting behavior is also described there. This section is self-contained and may be skipped by the reader uninterested in mathematical details. Finally, in the subsequent section, we carry out a simulation study to compare the accuracy of various approximate CSDs and demonstrate that ours are generally the most accurate.  相似文献   

14.
Telomere binding proteins protect chromosome ends from degradation and mask chromosome termini from checkpoint surveillance. In Saccharomyces cerevisiae, Cdc13 binds single-stranded G-rich telomere repeats, maintaining telomere integrity and length. Two additional proteins, Ten1 and Stn1, interact with Cdc13 but their contributions to telomere integrity are not well defined. Ten1 is known to prevent accumulation of aberrant single-stranded telomere DNA; whether this results from defective end protection or defective telomere replication is unclear. Here we report our analysis of a new group of ten1 temperature-sensitive (ts) mutants. At permissive temperatures, ten1-ts strains display greatly elongated telomeres. After shift to nonpermissive conditions, however, ten1-ts mutants accumulate extensive telomeric single-stranded DNA. Cdk1 activity is required to generate these single-stranded regions, and deleting the EXO1 nuclease partially suppresses ten1-ts growth defects. This is similar to cdc13-1 mutants, suggesting ten1-ts strains are defective for end protection. Moreover, like Cdc13, our analysis reveals Ten1 promotes de novo telomere addition. Interestingly, in ten1-ts strains at high temperatures, telomeric single-stranded DNA and Rad52-YFP repair foci are strongly induced despite Cdc13 remaining associated with telomeres, revealing Cdc13 telomere binding is not sufficient for end protection. Finally, unlike cdc13-1 mutants, ten1-ts strains display strong synthetic interactions with mutations in the POLα complex. These results emphasize that Cdc13 relies on Ten1 to execute its essential function, but leave open the possibility that Ten1 has a Cdc13-independent role in DNA replication.GENOME stability is critically dependent upon functional telomeres. DNA ends that lack telomeres, or that have dysfunctional telomeres, are metabolized by DNA repair processes; without an appropriate repair template, such chromosome ends can be resected or joined inappropriately with other chromosome ends. Thus, genomic integrity can be significantly compromised by telomere dysfunction, particularly in proliferating cells where cycles of instability may ensue due to creation of dicentric chromosomes (Bailey and Murnane 2006). Protein complexes that bind to the duplex and single-stranded telomere repeats are key for stabilizing the chromosome ends (de Lange 2005). In proliferating cells, this job is complicated not only because the terminal chromatin must be opened during the process of chromosome replication, but also because additional processes that metabolize DNA ends are active. For example, while nonhomologous end joining processes are preferentially used in repair of DNA double-strand breaks in G1, homologous recombination is preferentially used for this repair in S and G2 (Ferreira and Cooper 2004; Zierhut and Diffley 2008). Given these complexities, it is not surprising that our molecular understanding of how telomere proteins protect chromosomes ends is incomplete.Budding yeast has been useful for dissecting how cells correctly metabolize their chromosome ends. In Saccharomyces cerevisiae, the terminal DNA comprises approximately 300 bp of TG1-3/C1-3A sequences, ending with a short single-stranded overhang of the G-rich repeats. This 3′ overhang is ∼12–14 nucleotides, although during the late S/G2 phase of the cell cycle, it becomes longer, >30 nucleotides in length (Wellinger et al. 1993b; Dionne and Wellinger 1996; Larrivee et al. 2004). Central among factors that prevent inappropriate telomere degradation in S. cerevisiae is Cdc13, a protein that binds to single-stranded telomere G-rich repeats (Garvik et al. 1995; Lin and Zakian 1996; Nugent et al. 1996). Reducing Cdc13 function through either the cdc13-1 temperature sensitive (ts) allele or the cdc13-td conditional null (degron) allele results in telomere C-strand loss, with degradation continuing into the subtelomeric chromosomal regions (Garvik et al. 1995; Vodenicharov and Wellinger 2006). Correspondingly, homologous recombination at chromosome termini increases in cdc13-1 strains (Carson and Hartwell 1985; Garvik et al. 1995). The loss of Cdc13 unmasks the telomeres, provoking activation of the DNA damage checkpoint (Weinert and Hartwell 1993; Garvik et al. 1995). This protective role of Cdc13 is most likely its essential function.A thorough, mechanistic understanding of how Cdc13 mediates chromosome end protection is hampered in part because the activities responsible for the loss of the telomere C strand are not fully known. At normal telomeres, the Mre11-Rad50-Xrs2 complex has a role regulating resection required for telomere addition, whereas the Exo1 nuclease, Rad9 and Rad24 checkpoint proteins each influence the resection process at uncapped telomeres (Lydall and Weinert 1995; Maringele and Lydall 2002; Larrivee et al. 2004; Zubko et al. 2004). The 5′-to-3′ resection of both normal and uncapped telomeres is regulated by the activity of Cdk1, the yeast cyclin-dependent kinase (Frank et al. 2006; Vodenicharov and Wellinger 2006). Similar to the activities that promote 5′-to-3′ degradation of DNA ends at double-strand breaks (Aylon et al. 2004; Ira et al. 2004), the activities that lead to telomere resection are active in late S and G2 cell cycle phases (Wellinger et al. 1993a, 1996; Marcand et al. 2000; Vodenicharov and Wellinger 2006). Interestingly, Cdc13 is required to prevent degradation at telomeres only in proliferating cells and not when cells are blocked in stationary phase (Vodenicharov and Wellinger 2006). Additional factors, such as the S. cerevisiae Rap1 protein, prevent chromosome fusions by nonhomologous recombination during the G1 phase of the cell cycle (Pardo and Marcand 2005; Marcand et al. 2008).At least two additional proteins, Stn1 and Ten1, aid the capping role of Cdc13. Like CDC13, both STN1 and TEN1 are essential, and loss of their function leads to excessive single-stranded telomeric DNA (Grandin et al. 1997, 2001; Petreaca et al. 2007). STN1 was originally identified as a high copy suppressor of cdc13-1 temperature sensitivity (Grandin et al. 1997), and TEN1 was similarly isolated as a dosage suppressor of stn1-13 (Grandin et al. 2001). Combining either the cdc13-1 allele with stn1 mutations or the ten1-31 allele with stn1-13 is lethal (Grandin et al. 2001; Petreaca et al. 2007). The essential nature of these genes makes it difficult to clearly differentiate whether these genes operate in the same, or in parallel pathways to protect telomeres. A compelling argument that Cdc13, Stn1, and Ten1 likely function in a common pathway is that, in addition to these genetic interactions, Stn1 and Ten1 proteins interact with one another both in vivo and in vitro (Grandin et al. 2001; Gao et al. 2007), and each associates with Cdc13 in the yeast two-hybrid assay (Grandin et al. 1997, 2001; Petreaca et al. 2007). From these data, Cdc13, Stn1, and Ten1 are suggested to function as a single complex that mediates chromosome end protection in S. cerevisiae. Such a complex would share some similarities with the single-stranded DNA binding complex RPA (Gao et al. 2007). Whether these proteins normally operate exclusively as a heterotrimeric complex is still not entirely clear. Stn1 and Ten1 can make contributions to capping that are independent of Cdc13, as shown in experiments where overproducing the Stn1 essential domain with Ten1 replaced the essential function of Cdc13 (Petreaca et al. 2006). In addition, while the Schizosaccharomyces pombe Stn1 and Ten1 homologs are critical for telomere protection, they do not interact with Pot1, the single-stranded telomere binding protein that is also critical for telomere capping (Martin et al. 2007).The role of Ten1 in maintaining both telomere integrity and length homeostasis is not understood. It has been assumed that Stn1 and Ten1 play the same role as Cdc13 in maintaining telomere integrity, namely, preventing inappropriate terminal resection. However, whether this is in fact the case is not entirely clear. For one, disrupting the DNA replication machinery can give rise to an excess of terminal single-stranded DNA, although in this case, the ssDNA accumulation is attributed to a failure to synthesize the lagging DNA strand rather than removing a block to telomere resection (Diede and Gottschling 1999; Adams Martin et al. 2000). Although both Cdc13 and Stn1 are thought to act as capping proteins, each can interact with Polα subunits (Qi et al. 2003; Grossi et al. 2004; Petreaca et al. 2006), making it important to evaluate Ten1 function more carefully. Our goal here was to compare how Cdc13 and Ten1 promote chromosome end protection, first by testing whether Ten1 acts to prevent telomere resection from activities comparable to those that degrade telomeres in cdc13-1, and second by determining the impact of ten1 dysfunction upon Cdc13. The cdc13-1 allele has been extremely useful in analyzing the CDC13 essential function; TEN1 analysis has been hindered by a lack of equivalent genetic reagents. Here we have created a collection of ten1-ts alleles useful for probing the essential role of TEN1. Analysis of these alleles, which show constitutive telomere elongation, reveals that Ten1 promotes telomere capping with a similar cell cycle dependency as Cdc13, protecting ends during the period in which mitotic forms of Cdk1 are active. Critically, by showing that single-stranded DNA is generated in ten1-ts strains under conditions where semi-conservative replication is complete, we conclude that Ten1 truly can function as a capping protein. Moreover, the ten1-ts strains fail to restrain degradation of chromosome ends and induce formation of Rad52 repair foci, despite the association of wild-type Cdc13 with telomeres, indicating not only that Cdc13 binds telomeres independent of Ten1 function, but also that Cdc13 telomere localization is not sufficient for end protection. Finally, although the ten1-ts capping-deficient phenotypes parallel cdc13-1, only the ten1-ts strains are highly sensitive to impaired POL1 function, leaving open the possibility that TEN1 function additionally impacts terminal replication.  相似文献   

15.
A major question about cytokinesis concerns the role of the septin proteins, which localize to the division site in all animal and fungal cells but are essential for cytokinesis only in some cell types. For example, in Schizosaccharomyces pombe, four septins localize to the division site, but deletion of the four genes produces only a modest delay in cell separation. To ask if the S. pombe septins function redundantly in cytokinesis, we conducted a synthetic-lethal screen in a septin-deficient strain and identified seven mutations. One mutation affects Cdc4, a myosin light chain that is an essential component of the cytokinetic actomyosin ring. Five others cause frequent cell lysis during cell separation and map to two loci. These mutations and their dosage suppressors define a signaling pathway (including Rho1 and a novel arrestin) for repairing cell-wall damage. The seventh mutation affects the poorly understood RNA-binding protein Scw1 and severely delays cell separation when combined either with a septin mutation or with a mutation affecting the septin-interacting, anillin-like protein Mid2, suggesting that Scw1 functions in a pathway parallel to that of the septins. Taken together, our results suggest that the S. pombe septins participate redundantly in one or more pathways that cooperate with the actomyosin ring during cytokinesis and that a septin defect causes septum defects that can be repaired effectively only when the cell-integrity pathway is intact.THE fission yeast Schizosaccharomyces pombe provides an outstanding model system for studies of cytokinesis (McCollum and Gould 2001; Balasubramanian et al. 2004; Pollard and Wu 2010). As in most animal cells, successful cytokinesis in S. pombe requires an actomyosin ring (AMR). The AMR begins to assemble at the G2/M transition and involves the type II myosin heavy chains Myo2 and Myp2 and the light chains Cdc4 and Rlc1 (Wu et al. 2003). Myo2 and Cdc4 are essential for cytokinesis under all known conditions, Rlc1 is important at all temperatures but essential only at low temperatures, and Myp2 is essential only under stress conditions. As the AMR constricts, a septum of cell wall is formed between the daughter cells. The primary septum is sandwiched by secondary septa and subsequently digested to allow cell separation (Humbel et al. 2001; Sipiczki 2007). Because of the internal turgor pressure of the cells, the proper assembly and structural integrity of the septal layers are essential for cell survival.Septum formation involves the β-glucan synthases Bgs1/Cps1/Drc1, Bgs3, and Bgs4 (Ishiguro et al. 1997; Le Goff et al. 1999; Liu et al. 1999, 2002; Martín et al. 2003; Cortés et al. 2005) and the α-glucan synthase Ags1/Mok1 (Hochstenbach et al. 1998; Katayama et al. 1999). These synthases are regulated by the Rho GTPases Rho1 and Rho2 and the protein kinase C isoforms Pck1 and Pck2 (Arellano et al. 1996, 1997, 1999; Nakano et al. 1997; Hirata et al. 1998; Calonge et al. 2000; Sayers et al. 2000; Ma et al. 2006; Barba et al. 2008; García et al. 2009b). The Rho GTPases themselves appear to be regulated by both GTPase-activating proteins (GAPs) and guanine-nucleotide-exchange factors (GEFs) (Nakano et al. 2001; Calonge et al. 2003; Iwaki et al. 2003; Tajadura et al. 2004; Morrell-Falvey et al. 2005; Mutoh et al. 2005; García et al. 2006, 2009a,b). In addition, septum formation and AMR function appear to be interdependent. In the absence of a normal AMR, cells form aberrant septa and/or deposit septal materials at random locations, whereas a mutant defective in septum formation (bgs1) is also defective in AMR constriction (Gould and Simanis 1997; Le Goff et al. 1999; Liu et al. 1999, 2000). Both AMR constriction and septum formation also depend on the septation initiation network involving the small GTPase Spg1 (McCollum and Gould 2001; Krapp and Simanis 2008). Despite this considerable progress, many questions remain about the mechanisms and regulation of septum formation and its relationships to the function of the AMR.One major question concerns the role(s) of the septins. Proteins of this family are ubiquitous in fungal and animal cells and typically localize to the cell cortex, where they appear to serve as scaffolds and diffusion barriers for other proteins that participate in a wide variety of cellular processes (Longtine et al. 1996; Gladfelter et al. 2001; Hall et al. 2008; Caudron and Barral 2009). Despite the recent progress in elucidating the mechanisms of septin assembly (John et al. 2007; Sirajuddin et al. 2007; Bertin et al. 2008; McMurray and Thorner 2008), the details of septin function remain obscure. However, one prominent role of the septins and associated proteins is in cytokinesis. Septins concentrate at the division site in every cell type that has been examined, and in Saccharomyces cerevisiae (Hartwell 1971; Longtine et al. 1996; Lippincott et al. 2001; Dobbelaere and Barral 2004) and at least some Drosophila (Neufeld and Rubin 1994; Adam et al. 2000) and mammalian (Kinoshita et al. 1997; Surka et al. 2002) cell types, the septins are essential for cytokinesis. In S. cerevisiae, the septins are required for formation of the AMR (Bi et al. 1998; Lippincott and Li 1998). However, this cannot be their only role, because the AMR itself is not essential for cytokinesis in this organism (Bi et al. 1998; Korinek et al. 2000; Schmidt et al. 2002). Moreover, there is no evidence that the septins are necessary for AMR formation or function in any other organism. A further complication is that in some cell types, including most Caenorhabditis elegans cells (Nguyen et al. 2000; Maddox et al. 2007) and some Drosophila cells (Adam et al. 2000; Field et al. 2008), the septins do not appear to be essential for cytokinesis even though they localize to the division site.S. pombe has seven septins, four of which (Spn1, Spn2, Spn3, and Spn4) are expressed in vegetative cells and localize to the division site shortly before AMR constriction and septum formation (Longtine et al. 1996; Berlin et al. 2003; Tasto et al. 2003; Wu et al. 2003; An et al. 2004; Petit et al. 2005; Pan et al. 2007; Onishi et al. 2010). Spn1 and Spn4 appear to be the core members of the septin complex (An et al. 2004; McMurray and Thorner 2008), and mutants lacking either of these proteins do not assemble the others at the division site. Assembly of a normal septin ring also depends on the anillin-like protein Mid2, which colocalizes with the septins (Berlin et al. 2003; Tasto et al. 2003). Surprisingly, mutants lacking the septins are viable and form seemingly complete septa with approximately normal timing. These mutants do, however, display a variable delay in separation of the daughter cells, suggesting that the septins play some role(s) in the proper completion of the septum or in subsequent processes necessary for cell separation (Longtine et al. 1996; An et al. 2004; Martín-Cuadrado et al. 2005).It is possible that the septins localize to the division site and yet are nonessential for division in some cell types because their role is redundant with that of some other protein(s) or pathway(s). To explore this possibility in S. pombe, we screened for mutations that were lethal in combination with a lack of septins. The results suggest that the septins cooperate with the AMR during cytokinesis and that, in the absence of septin function, the septum is not formed properly, so that an intact system for recognizing and repairing cell-wall damage becomes critical for cell survival.  相似文献   

16.
17.
Effective population size (Ne) is a central evolutionary concept, but its genetic estimation can be significantly complicated by age structure. Here we investigate Ne in Atlantic salmon (Salmo salar) populations that have undergone changes in demography and population dynamics, applying four different genetic estimators. For this purpose we use genetic data (14 microsatellite markers) from archived scale samples collected between 1951 and 2004. Through life table simulations we assess the genetic consequences of life history variation on Ne. Although variation in reproductive contribution by mature parr affects age structure, we find that its effect on Ne estimation may be relatively minor. A comparison of estimator models suggests that even low iteroparity may upwardly bias Ne estimates when ignored (semelparity assumed) and should thus empirically be accounted for. Our results indicate that Ne may have changed over time in relatively small populations, but otherwise remained stable. Our ability to detect changes in Ne in larger populations was, however, likely hindered by sampling limitations. An evaluation of Ne estimates in a demographic context suggests that life history diversity, density-dependent factors, and metapopulation dynamics may all affect the genetic stability of these populations.THE effective size of a population (Ne) is an evolutionary parameter that can be informative on the strength of stochastic evolutionary processes, the relevance of which relative to deterministic forces has been debated for decades (e.g., Lande 1988). Stochastic forces include environmental, demographic, and genetic components, the latter two of which are thought to be more prominent at reduced population size, with potentially detrimental consequences for average individual fitness and population persistence (Newman and Pilson 1997; Saccheri et al. 1998; Frankham 2005). The quantification of Ne in conservation programs is thus frequently advocated (e.g., Luikart and Cornuet 1998; Schwartz et al. 2007), although gene flow deserves equal consideration given its countering effects on genetic stochasticity (Frankham et al. 2003; Palstra and Ruzzante 2008).Effective population size is determined mainly by the lifetime reproductive success of individuals in a population (Wright 1938; Felsenstein 1971). Variance in reproductive success, sex ratio, and population size fluctuations can reduce Ne below census population size (Frankham 1995). Given the difficulty in directly estimating Ne through quantification of these demographic factors (reviewed by Caballero 1994), efforts have been directed at inferring Ne indirectly through measurement of its genetic consequences (see Leberg 2005, Wang 2005, and Palstra and Ruzzante 2008 for reviews). Studies employing this approach have quantified historical levels of genetic diversity and genetic threats to population persistence (e.g., Nielsen et al. 1999b; Miller and Waits 2003; Johnson et al. 2004). Ne has been extensively studied in (commercially important) fish species, due to the common availability of collections of archived samples that facilitate genetic estimation using the temporal method (e.g., Hauser et al. 2002; Shrimpton and Heath 2003; Gomez-Uchida and Banks 2006; Saillant and Gold 2006).Most models relating Ne to a population''s genetic behavior make simplifying assumptions regarding population dynamics. Chiefly among these is the assumption of discrete generations, frequently violated in practice given that most natural populations are age structured with overlapping generations. Here, theoretical predictions still apply, provided that population size and age structure are constant (Felsenstein 1971; Hill 1972). Ignored age structure can introduce bias into temporal genetic methods for the estimation of Ne, especially for samples separated by time spans that are short relative to generation interval (Jorde and Ryman 1995; Waples and Yokota 2007; Palstra and Ruzzante 2008). Moreover, estimation methods that do account for age structure (e.g., Jorde and Ryman 1995) still assume this structure to be constant. Population dynamics will, however, likely be altered as population size changes, thus making precise quantifications of the genetic consequences of acute population declines difficult (Nunney 1993; Engen et al. 2005; Waples and Yokota 2007). This problem may be particularly relevant when declines are driven by anthropogenic impacts, such as selective harvesting regimes, that can affect age structure and Ne simultaneously (Ryman et al. 1981; Allendorf et al. 2008). Demographic changes thus have broad conservation implications, as they can affect a population''s sensitivity to environmental stochasticity and years of poor recruitment (Warner and Chesson 1985; Ellner and Hairston 1994; Gaggiotti and Vetter 1999). Consequently, although there is an urgent need to elucidate the genetic consequences of population declines, relatively little is understood about the behavior of Ne when population dynamics change (but see Engen et al. 2005, 2007).Here we focus on age structure and Ne in Atlantic salmon (Salmo salar) river populations in Newfoundland and Labrador. The freshwater habitat in this part of the species'' distribution range is relatively pristine (Parrish et al. 1998), yet Atlantic salmon in this area have experienced demographic declines, associated with a commercial marine fishery, characterized by high exploitation rates (40–80% of anadromous runs; Dempson et al. 2001). A fishery moratorium was declared in 1992, with rivers displaying differential recovery patterns since then (Dempson et al. 2004b), suggesting a geographically variable impact of deterministic and stochastic factors, possibly including genetics. An evaluation of those genetic consequences thus requires accounting for potential changes in population dynamics as well as in life history. Life history in Atlantic salmon can be highly versatile (Fleming 1996; Hutchings and Jones 1998; Fleming and Reynolds 2004), as exemplified by the high variation in age-at-maturity displayed among and within populations (Hutchings and Jones 1998), partly reflecting high phenotypic plasticity (Hutchings 2004). This diversity is particularly evident in the reproductive biology of males, which can mature as parr during juvenile freshwater stages (Jones and King 1952; Fleming and Reynolds 2004) and/or at various ages as anadromous individuals, when returning to spawn in freshwater from ocean migration. Variability in life history strategies is further augmented by iteroparity, which can be viewed as a bet-hedging strategy to deal with environmental uncertainty (e.g., Orzack and Tuljapurkar 1989; Fleming and Reynolds 2004). Life history diversity and plasticity may allow salmonid fish populations to alter and optimize their life history under changing demography and population dynamics, potentially acting to stabilize Ne. Reduced variance in individual reproductive success at low breeder abundance (genetic compensation) will achieve similar effects and might be a realistic aspect of salmonid breeding systems (Ardren and Kapuscinski 2003; Fraser et al. 2007b). Little is currently known about the relationships between life history plasticity, demographic change and Ne, partly due to scarcity of the multivariate data required for these analyses.Our objective in this article is twofold. First, we use demographic data for rivers in Newfoundland to quantify how life history variation influences age structure in Atlantic salmon and hence Ne and its empirical estimation from genetic data. We find that variation in reproductive contribution by mature parr has a much smaller effect on the estimation of Ne than is often assumed. Second, we use temporal genetic data to estimate Ne and quantify the genetic consequences of demographic changes. We attempt to account for potential sources of bias, associated with (changes in) age structure and life history, by using four different analytical models to estimate Ne: a single-sample estimator using the linkage disequilibrium method (Hill 1981), the temporal model assuming discrete generations (Nei and Tajima 1981; Waples 1989), and two temporal models for species with overlapping generations (Waples 1990a,b; Jorde and Ryman 1995) that differ principally in assumptions regarding iteroparity. A comparison of results from these different estimators suggests that iteroparity may often warrant analytical consideration, even when it is presumably low. Although sometimes limited by statistical power, a quantification and comparison of temporal changes in Ne among river populations suggests a more prominent impact of demographic changes on Ne in relatively small river populations.  相似文献   

18.
19.
20.
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