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.
The intraflagellar transport machinery is required for the assembly of cilia. It has been investigated by biochemical, genetic, and computational methods that have identified at least 21 proteins that assemble into two subcomplexes. It has been hypothesized that complex A is required for retrograde transport. Temperature-sensitive mutations in FLA15 and FLA17 show defects in retrograde intraflagellar transport (IFT) in Chlamydomonas. We show that IFT144 and IFT139, two complex A proteins, are encoded by FLA15 and FLA17, respectively. The fla15 allele is a missense mutation in a conserved cysteine and the fla17 allele is an in-frame deletion of three exons. The flagellar assembly defect of each mutant is rescued by the respective transgenes. In fla15 and fla17 mutants, bulges form in the distal one-third of the flagella at the permissive temperature and this phenotype is also rescued by the transgenes. These bulges contain the complex B component IFT74/72, but not α-tubulin or p28, a component of an inner dynein arm, which suggests specificity with respect to the proteins that accumulate in these bulges. IFT144 and IFT139 are likely to interact with each other and other proteins on the basis of three distinct genetic tests: (1) Double mutants display synthetic flagellar assembly defects at the permissive temperature, (2) heterozygous diploid strains exhibit second-site noncomplemention, and (3) transgenes confer two-copy suppression. Since these tests show different levels of phenotypic sensitivity, we propose they illustrate different gradations of gene interaction between complex A proteins themselves and with a complex B protein (IFT172).CILIA and flagella are microtubule-based organelles that are found on most mammalian cells. They provide motility to cells and participate in many sensory processes. Defects in or loss of cilia/flagella cause a variety of human diseases that include polycystic kidney disease, retinal degeneration, infertility, obesity, respiratory defects, left–right axis determination, and polydactyly (Fliegauf et al. 2007). Mouse mutants demonstrate that cilia are essential for viability, neural tube closure, and bone development (Eggenschwiler and Anderson 2007; Fliegauf et al. 2007). Cilia and flagella are also present in protists, algae, moss, and some fungi.The assembly and maintenance of cilia and flagella require intraflagellar transport (IFT) (Kozminski et al. 1995). IFT involves the movement of 100- to 200-nm-long protein particles from the basal body located in the cell body to the tip of the flagella using the heterotrimeric kinesin-2 (anterograde movement) (Kozminski et al. 1995) and movement back to the cell body (retrograde movement) using the cytoplasmic dynein complex (Pazour et al. 1999; Porter et al. 1999). IFT particles change their direction of movement as well as their size, speed, and frequency at the ends of the flagella as they switch from anterograde to retrograde movement (Iomini et al. 2001). Biochemical isolation of IFT particles reveals that they are composed of at least 16 proteins and that these particles can be dissociated into two complexes in vitro by changing the salt concentration (Cole et al. 1998; Piperno et al. 1998). Recent genetic and bioinformatics analysis adds at least 7 more proteins to the IFT particle (Follit et al. 2009) (Eggenschwiler and Anderson 2007).

TABLE 1

Proteins and gene names for the intraflagellar transport particles in Chlamydomonas, C. elegans, and mouse
ProteinMotifChlamydomonas geneC. elegans geneMouse geneReferences to worm and mouse genes
Complex A
IFT144WDFLA15
IFT140WDche-11Qin et al. (2001)
IFT139TRPFLA17dyf-2THM1Efimenko et al. (2006); Tran et al. (2008)
IFT122WDIFTA-1Blacque et al. (2006)
IFT121WDdaf-10Bell et al. (2006)
IFT43
Complex B
IFT172WDFLA11osm-1WimpleHuangfu et al. (2003); Pedersen et al. (2005); Bell et al. (2006)
IFT88TRPIFT88osm-5Tg737/PolarisPazour et al. (2000); Qin et al. (2001)
IFT81Coilift-81CDV1Kobayashi et al. (2007)
IFT80WDche-2Wdr56Fujiwara et al. (1999)
IFT74/72Coilift-74Cmg1Kobayashi et al. (2007)
IFT57/55Coilche-13HippiHaycraft et al. (2003)
IFT54Microtubule binding domain MIP-T3dyf-11Traf3IP1Kunitomo and Iino (2008); Li et al. (2008); Omori et al. (2008); Follit et al. (2009)
IFT52ABC typeBLD1osm-6Ngd2Brazelton et al. (2001); Bell et al. (2006)
IFT46IFT46dyf-6Bell et al. (2006); Hou et al. (2007)
IFT27G proteinNot presentRabl4
IFT25Hsp20Not presentHSP16.1Follit et al. (2009)
IFT22G proteinIFTA-2Rabl5Schafer et al. (2006)
IFT20CoilFollit et al. (2006)
FAP22Cluamp related proteindyf-3Cluamp1Murayama et al. (2005); Follit et al. (2009)
DYF13


dyf-13
Ttc26
Blacque et al. (2005)
Open in a separate window—, no mutant found to date in Chlamydomonas.A collection of temperature-sensitive mutant strains that fail to assemble flagella at the restrictive temperature of 32° was isolated in Chlamydomonas (Huang et al. 1977; Adams et al. 1982; Piperno et al. 1998; Iomini et al. 2001). Analysis of the flagella at 21° permits the measurement of the velocity and frequency of IFT particles in the mutant strains. This analysis suggested that assembly has four phases: recruitment to the basal body, anterograde movement (phases I and II), retrograde movement, and return to the cytoplasm (phases III and IV) (Iomini et al. 2001). Different mutants were classified as defective in these four phases. However, because different alleles of FLA8 were classified as defective in different phases (Iomini et al. 2001; Miller et al. 2005), we combined mutants with IFT defects into just two classes. The first group (phases I and II) includes mutant strains that show decreased anterograde velocities, a decreased ratio of anterograde to retrograde particles, and an accumulation of complex A proteins at the basal body. This group includes mutations in the FLA8 and FLA10 genes, which encode the two motor subunits of kinesin-2 (Walther et al. 1994; Miller et al. 2005), as well as mutations in three unknown genes (FLA18, FLA27, and FLA28). The second group includes mutant strains that show the reciprocal phenotype (phases III and IV); these phenotypes include decreased retrograde velocities, an increased ratio of anterograde to retrograde particles, and an accumulation of complex B proteins in the flagella. With the exception of the FLA11 gene, which encodes IFT172, a component of complex B (Pedersen et al. 2005), the gene products in this class are unknown (FLA2, FLA15, FLA16, FLA17, and FLA24). One might predict that mutations in this group would map to genes that encode complex A or retrograde motor subunits. Interestingly, IFT particles isolated from fla11, fla15, fla16, and fla17-1 flagella show depletion of complex A polypeptides (Piperno et al. 1998; Iomini et al. 2001). The inclusion of IFT172 in this class is explained by the observations that IFT172 plays a role in remodeling the IFT particles at the flagellar tip to transition from anterograde to retrograde movement (Pedersen et al. 2005). The remaining mutant strains do not show obvious defects in velocities, ratios, or accumulation at 21° and may reflect a less severe phenotype at the permissive temperature or a non-IFT role for these genes.Direct interactions occur between components of complex B. IFT81 and IFT74/72 interact to form a scaffold required for IFT complex B assembly (Lucker et al. 2005). IFT57 and IFT20 also interact with each other and kinesin-2 (Baker et al. 2003). While physical interactions are being used to define IFT particle architecture, genetic interactions among loci encoding IFT components should be instructive regarding their function as well. To probe retrograde movement and its function, we have identified the gene products encoded by two retrograde defective mutant strains. They are FLA15 and FLA17 and encode IFT144 and IFT139, respectively. The genetic interactions of these loci provide interesting clues about the assembly of the IFT particles and possible physical interactions in the IFT particles.  相似文献   

2.
3.
4.
Insertional mutagenesis screens play an integral part in the annotating of functional data for all sequenced genes in the postgenomic era. Chemical mutagenesis screens are highly efficient but identifying the causative gene can be a laborious task. Other mutagenesis platforms, such as transposable elements, have been successfully applied for insertional mutagenesis screens in both the mouse and rat. However, relatively low transposition efficiency has hampered their use as a high-throughput forward genetic mutagenesis screen. Here we report the first evidence of germline activity in the mouse using a naturally active DNA transposon derived from the medaka fish called Tol2, as an alternative system for high-throughput forward genetic mutagenesis screening tool.THE Tol2 (transposable element of Oryzias latipes, number 2) element belongs to the hAT family (hobo of Drosophilia, Activator of maize and Tam3 of snapdragon) of transposons and was the first known autonomously active vertebrate type II transposable element (Koga et al. 1996; Kawakami et al. 1998). Unlike other DNA-type transposons like Sleeping Beauty (SB) (Ivics et al. 1997) or piggyBac (PB) (Fraser et al. 1996), Tol2 does not exhibit any known strong site specificity for integration nor does it exhibit any significant overexpression inhibition activity (Kawakami and Noda 2004; Balciunas et al. 2006) as seen in SB (Geurts et al. 2003). Recently, Tol2 was shown to effectively carry large DNA cargo of up to 10 kb in human and mouse cells without affecting its transposition efficiency (Balciunas et al. 2006). To date, Tol2 has also been demonstrated to transpose efficiently in zebrafish, frog, chicken, mouse cells, and human cells (Kawakami et al. 2000, 2004; Koga et al. 2003; Kawakami and Noda 2004; Balciunas et al. 2006; Hamlet et al. 2006; Sato et al. 2007).Germline mutagenesis using the SB transposon system has been demonstrated in both the mouse (Dupuy et al. 2001; Horie et al. 2001) and rat (Kitada et al. 2007; Lu et al. 2007). In addition, PB germline mutagenesis in mice has also been demonstrated (Ding et al. 2005; Wu et al. 2007). However, the relatively low germline transposition efficiency of both transposon systems reported so far has hampered their use in a high-throughput forward genetic mutagenesis screen (Keng et al. 2005; Kitada et al. 2007).

TABLE 1

Germline transposition frequency in various transposon systems
Transposon systemAverage transposition events per gameteMouse strainReference
SB2FVB/NDupuy et al. (2001)
SB1.25C3H and C57BL/6Horie et al. (2001)
SB1.15C3H and C57BL/6Keng et al. (2005)
PB1.1FVB/NDing et al. (2005)
PB1C57BL/6Wu et al. (2007)
Tol2
3
FVB/N
Present study
Open in a separate windowSB, Sleeping Beauty; PB, piggyBac; Tol2, transposable element of Oryzias latipes, number 2.In search of an alternative tool for high-throughput forward germline mutagenesis screen in mice, a Tol2 transposon insertional mutagenesis system was generated on the basis of a similar strategy used for the SB transposon system (Horie et al. 2003; Keng et al. 2005). In the present study, we successfully demonstrate the novel use of the Tol2 transposon system for germline mutagenesis in mouse. Our results indicate the potential use of this transposon system for a high-throughput, large-scale forward mutagenesis screen in the mouse germline.  相似文献   

5.
The capacity for phenotypic evolution is dependent upon complex webs of functional interactions that connect genotype and phenotype. Wrinkly spreader (WS) genotypes arise repeatedly during the course of a model Pseudomonas adaptive radiation. Previous work showed that the evolution of WS variation was explained in part by spontaneous mutations in wspF, a component of the Wsp-signaling module, but also drew attention to the existence of unknown mutational causes. Here, we identify two new mutational pathways (Aws and Mws) that allow realization of the WS phenotype: in common with the Wsp module these pathways contain a di-guanylate cyclase-encoding gene subject to negative regulation. Together, mutations in the Wsp, Aws, and Mws regulatory modules account for the spectrum of WS phenotype-generating mutations found among a collection of 26 spontaneously arising WS genotypes obtained from independent adaptive radiations. Despite a large number of potential mutational pathways, the repeated discovery of mutations in a small number of loci (parallel evolution) prompted the construction of an ancestral genotype devoid of known (Wsp, Aws, and Mws) regulatory modules to see whether the types derived from this genotype could converge upon the WS phenotype via a novel route. Such types—with equivalent fitness effects—did emerge, although they took significantly longer to do so. Together our data provide an explanation for why WS evolution follows a limited number of mutational pathways and show how genetic architecture can bias the molecular variation presented to selection.UNDERSTANDING—and importantly, predicting—phenotypic evolution requires knowledge of the factors that affect the translation of mutation into phenotypic variation—the raw material of adaptive evolution. While much is known about mutation rate (e.g., Drake et al. 1998; Hudson et al. 2002), knowledge of the processes affecting the translation of DNA sequence variation into phenotypic variation is minimal.Advances in knowledge on at least two fronts suggest that progress in understanding the rules governing the generation of phenotypic variation is possible (Stern and Orgogozo 2009). The first stems from increased awareness of the genetic architecture underlying specific adaptive phenotypes and recognition of the fact that the capacity for evolutionary change is likely to be constrained by this architecture (Schlichting and Murren 2004; Hansen 2006). The second is the growing number of reports of parallel evolution (e.g., Pigeon et al. 1997; ffrench-Constant et al. 1998; Allender et al. 2003; Colosimo et al. 2004; Zhong et al. 2004; Boughman et al. 2005; Shindo et al. 2005; Kronforst et al. 2006; Woods et al. 2006; Zhang 2006; Bantinaki et al. 2007; McGregor et al. 2007; Ostrowski et al. 2008)—that is, the independent evolution of similar or identical features in two or more lineages—which suggests the possibility that evolution may follow a limited number of pathways (Schluter 1996). Indeed, giving substance to this idea are studies that show that mutations underlying parallel phenotypic evolution are nonrandomly distributed and typically clustered in homologous genes (Stern and Orgogozo 2008).While the nonrandom distribution of mutations during parallel genetic evolution may reflect constraints due to genetic architecture, some have argued that the primary cause is strong selection (e.g., Wichman et al. 1999; Woods et al. 2006). A means of disentangling the roles of population processes (selection) from genetic architecture is necessary for progress (Maynard Smith et al. 1985; Brakefield 2006); also necessary is insight into precisely how genetic architecture might bias the production of mutations presented to selection.Despite their relative simplicity, microbial populations offer opportunities to advance knowledge. The wrinkly spreader (WS) morphotype is one of many different niche specialist genotypes that emerge when experimental populations of Pseudomonas fluorescens are propagated in spatially structured microcosms (Rainey and Travisano 1998). Previous studies defined, via gene inactivation, the essential phenotypic and genetic traits that define a single WS genotype known as LSWS (Spiers et al. 2002, 2003) (Figure 1). LSWS differs from the ancestral SM genotype by a single nonsynonymous nucleotide change in wspF. Functionally (see Figure 2), WspF is a methyl esterase and negative regulator of the WspR di-guanylate cyclase (DGC) (Goymer et al. 2006) that is responsible for the biosynthesis of c-di-GMP (Malone et al. 2007), the allosteric activator of cellulose synthesis enzymes (Ross et al. 1987). The net effect of the wspF mutation is to promote physiological changes that lead to the formation of a microbial mat at the air–liquid interface of static broth microcosms (Rainey and Rainey 2003).Open in a separate windowFigure 1.—Outline of experimental strategy for elucidation of WS-generating mutations and their subsequent identity and distribution among a collection of independently evolved, spontaneously arising WS genotypes. The strategy involves, first, the genetic analysis of a specific WS genotype (e.g., LSWS) to identify the causal mutation, and second, a survey of DNA sequence variation at specific loci known to harbor causal mutations among a collection of spontaneously arising WS genotypes. For example, suppressor analysis of LSWS using a transposon to inactivate genes necessary for expression of the wrinkly morphology delivered a large number of candidate genes (top left) (Spiers et al. 2002). Genetic and functional analysis of these candidate genes (e.g., Goymer et al. 2006) led eventually to the identity of the spontaneous mutation (in wspF) responsible for the evolution of LSWS from the ancestral SM genotype (Bantinaki et al. 2007). Subsequent analysis of the wspF sequence among 26 independent WS genotypes (bottom) showed that 50% harbored spontaneous mutations (of different kinds; see Open in a separate windowFigure 2.—Network diagram of DGC-encoding pathways underpinning the evolution of the WS phenotype and their regulation. Overproduction of c-di-GMP results in overproduction of cellulose and other adhesive factors that determine the WS phenotype. The ancestral SBW25 genome contains 39 putative DGCs, each in principle capable of synthesizing the production of c-di-GMP, and yet WS genotypes arise most commonly as a consequence of mutations in just three DGC-containing pathways: Wsp, Aws, and Mws. In each instance, the causal mutations are most commonly in the negative regulatory component: wspF, awsX, and the phosphodiesterase domain of mwsR (see text).To determine whether spontaneous mutations in wspF are a common cause of the WS phenotype, the nucleotide sequence of this gene was obtained from a collection of 26 spontaneously arising WS genotypes (WSA-Z) taken from 26 independent adaptive radiations, each founded by the same ancestral SM genotype (Figure 1): 13 contained mutations in wspF (Bantinaki et al. 2007). The existence of additional mutational pathways to WS provided the initial motivation for this study.

TABLE 1

Mutational causes of WS
WS genotypeGeneNucleotide changeAmino acid changeSource/reference
LSWSwspFA901CS301RBantinaki et al. (2007)
AWSawsXΔ100-138ΔPDPADLADQRAQAThis study
MWSmwsRG3247AE1083KThis study
WSAwspFT14GI5SBantinaki et al. (2007)
WSBwspFΔ620-674P206Δ (8)aBantinaki et al. (2007)
WSCwspFG823TG275CBantinaki et al. (2007)
WSDwspEA1916GD638GThis study
WSEwspFG658TV220LBantinaki et al. (2007)
WSFwspFC821TT274IBantinaki et al. (2007)
WSGwspFC556TH186YBantinaki et al. (2007)
WSHwspEA2202CK734NThis study
WSIwspEG1915TD638YThis study
WSJwspFΔ865-868R288Δ (3)aBantinaki et al. (2007)
WSKawsOG125TG41VThis study
WSLwspFG482AG161DBantinaki et al. (2007)
WSMawsRC164TS54FThis study
WSNwspFA901CS301RBantinaki et al. (2007)
WSOwspFΔ235-249V79Δ (6)aBantinaki et al. (2007)
WSPawsR222insGCCACCGAA74insATEThis study
WSQmwsR3270insGACGTG1089insDVThis study
WSRmwsRT2183CV272AThis study
WSSawsXC472TQ158STOPThis study
WSTawsXΔ229-261ΔYTDDLIKGTTQThis study
WSUwspFΔ823-824T274Δ (13)aBantinaki et al. (2007)
WSVawsXT74GL24RThis study
WSWwspFΔ149L49Δ (1)aBantinaki et al. (2007)
WSXb???This study
WSYwspFΔ166-180Δ(L51-I55)Bantinaki et al. (2007)
WSZ
mwsR
G3055A
A1018T
This study
Open in a separate windowaP206Δ(8) indicates a frameshift; the number of new residues before a stop codon is reached is in parentheses.bSuppressor analysis implicates the wsp locus (17 transposon insertions were found in this locus). However, repeated sequencing failed to identify a mutation.Here we define and characterize two new mutational routes (Aws and Mws) that together with the Wsp pathway account for the evolution of 26 spontaneously arising WS genotypes. Each pathway offers approximately equal opportunity for WS evolution; nonetheless, additional, less readily realized genetic routes producing WS genotypes with equivalent fitness effects exist. Together our data show that regulatory pathways with specific functionalities and interactions bias the molecular variation presented to selection.  相似文献   

6.
DNA sequence analysis and genetic mapping of loci from mating-type-specific chromosomes of the smut fungus Microbotryum violaceum demonstrated that the nonrecombining mating-type-specific region in this species comprises ∼25% (∼1 Mb) of the chromosome length. Divergence between homologous mating-type-linked genes in this region varies between 0 and 8.6%, resembling the evolutionary strata of vertebrate and plant sex chromosomes.EVOLUTION of mating types or sex-determining systems often involves the suppression of recombination around the primary sex-determining or mating-type-determining locus. In animals and plants, it is often an entire or almost entire chromosome (Y or W in male or female heterogametic species, respectively) that ceases to recombine with its homologous (X or Z) chromosome (Charlesworth and Charlesworth 2000; Charlesworth 2008). Self-incompatibility loci in plants are also thought to be located in regions of suppressed recombination (Charlesworth et al. 2005; Kamau and Charlesworth 2005; Kamau et al. 2007; Li et al. 2007; Yang et al. 2007). Regardless of the phylogenetic position of a species, such nonrecombining regions are known to follow similar evolutionary trajectories. The nonrecombining region on the sex-specific chromosome expands in several steps, forming evolutionary strata—regions of different X/Y (or Z/W) divergence (Lahn and Page 1999; Handley et al. 2004; Sandstedt and Tucker 2004; Nicolas et al. 2005)—and genes in the nonrecombining regions gradually accumulate deleterious mutations that eventually render them dysfunctional (Charlesworth and Charlesworth 2005; Charlesworth 2008).Fungal mating-type systems are very diverse, with the number of mating types varying from two to several hundred (Casselton 2002). Like sex chromosomes in several animals and plants, suppressed recombination has evolved in regions near fungal mating-type loci, including in Ustilago hordei (Lee et al. 1999), Cryptococcus neoformans (Lengeler et al. 2002), and Neurospora tetrasperma (Menkis et al. 2008). These species have two mating types, but no morphologically distinct sexes. The mating-type locus (the region of suppressed recombination) of C. neoformans is small (∼100 kb) compared with known sex chromosomes and contains only ∼20 genes that, unlike many sex chromosomes (Y or W chromosomes), show no obvious signs of genetic degeneration (Lengeler et al. 2002; Fraser et al. 2004). Judging from the divergence between the homologous genes on the two mating-type-specific chromosomes, C. neoformans started to evolve sex chromosomes a long time ago because silent divergence between the two mating types in the most ancient region exceeds 100% (Fraser et al. 2004). Genes in the younger mating-type-specific region are much less diverged between the two sex chromosomes, suggesting that the evolution of the sex locus in C. neoformans might have proceeded through several steps. The nonrecombining region around the mating-type locus of N. tetrasperma is much larger than in C. neoformans (at least 6.6 Mb), and silent divergence between homologous genes on the mating-type-specific chromosomes ranges from zero to 9%, demonstrating that these mating-type-specific chromosomes evolved recently (Menkis et al. 2008).M. violaceum, which causes anther smut disease in Silene latifolia and other species in the family Caryophyllaceae, has two mating types, A1 and A2 (reviewed by Giraud et al. 2008), which are determined by the presence of mating-type-specific chromosomes (hereafter A1 and A2 chromosomes, or sex chromosomes) in the haploid stage of the life cycle (Hood 2002; Hood et al. 2004). The A1 and A2 chromosomes are distinguishable by size in pulsed-field electrophoresis, and it is possible to isolate individual chromosomes electrophoretically (Hood et al. 2004). Random fragments of A1 and A2 chromosomes have previously been isolated from mating-type-specific bands of pulsed-field separated chromosomes of M. violaceum (Hood et al. 2004). These fragments were assumed to be linked to mating type. The same method was used to isolate fragments of non-mating-type-specific chromosomes. On the basis of the analysis of their sequences, (Hood et al. 2004) proposed that mating-type-specific chromosomes in M. violaceum might be degenerate because they contained a lower proportion of protein-coding genes than other chromosomes. However, it was not determined whether the sequences isolated from the mating-type chromosomes originated from the mating-type-specific or from the recombining regions (Hood et al. 2004), and the relative sizes of these regions are not known for these M. violaceum chromosomes. We tested the mating-type specificity of 86 of these fragments and demonstrate that fewer than a quarter of these loci are located in the mating-type-specific region, suggesting that the nonrecombining region on the A1 and A2 chromosomes is quite small, while the rest of the chromosome probably recombines (like pseudoautosomal regions of sex chromosomes) and is therefore not expected to undergo genetic degeneration. Genetic mapping confirms the presence of two pseudoautosomal regions in the M. violaceum mating-type-specific chromosomes.As these chromosomes are mating type specific in the haploid stage of M. violaceum, mating-type-specific loci (or DNA fragments) can be identified by testing whether they are present exclusively in A1 or A2 haploid strains. We therefore prepared haploid A1 and A2 M. violaceum cultures from S. latifolia plants from two geographically remote locations (accessions Sl405 from Sweden and Sl127 from the French Pyrenees). Haploid sporidial cultures were isolated by a standard dilution method (Kaltz and Shykoff 1997; Oudemans and Alexander 1998). Mating types were determined by PCR amplification of each culture with primers designed for A1 and A2 pheromone receptor genes linked to A1 and A2 mating types (Yockteng et al. 2007). The primers were as follows: 5′-TGGCATCCCTCAATGTTTCC-3′ and 5′-CACCTTTTGATGAGAGGCCG-3′ for the A1 pheromone receptor (GenBank accession no. EF584742) and 5′-TGACGAGAGCATTCCTACCG-3′ and 5′-GAAGCGGAACTTGCCTTTCT-3′ for the A2 pheromone receptor (GenBank accession no. EF584741). Cultures with PCR product amplified only from an A1 or A2 pheromone receptor gene were selected for further use. The mating types of the cultures were verified by conjugating them in all combinations.The GenBank nucleotide database was searched using BLAST for sequences similar to those isolated by Hood et al. (2004). Sequences with similarity to transposable elements (TE) and other repeats were excluded. The resulting set of nonredundant sequences was used to design PCR primers for 98 fragments. Half of these were originally isolated from the A1 and half from the A2 chromosomes and are hereafter called A1-NNN or A2-NNN (where NNN is the locus number; supporting information, Table S1), which does not imply that these loci are A1 or A2 specific, but merely indicates that they were originally isolated from the A1 or A2 chromosomes. Amplification of these regions from new A1 and A2 M. violaceum cultures, independently isolated by ourselves, revealed that only 5 of the 49 loci isolated from the A1 chromosome are indeed A1 specific and only 6 of 49 isolated from the A2 chromosome are A2 specific. All other loci amplified from both A1 and A2 cultures. Figure 1 illustrates some of these results from the Swedish sample (Sl405).Open in a separate windowFigure 1.—Testing of mating-type specificity for loci isolated from A1 and A2 chromosomes. (a) PCR amplifications from haploid cultures from Sl405 using primers designed from six A1-originated loci. Loci in which a PCR product could be amplified only from A1 cultures (boxed) were classified as specific to mating type A1. (b) PCR tests of six A2-originated loci on the same set of haploids as in a. Loci in which a PCR product amplified only from A2 cultures (boxed) were classified as specific to mating type A2. Loci amplified from both A1 and A2 cultures are not mating type specific.The fragments that amplified from both A1 and A2 mating types may be in recombining regions, or they could be present in mating-type-specific regions on both A1 and A2 chromosomes. If they are in recombining regions, the A1- and A2-linked homologs should not be diverged from each other, but if they are in nonrecombining, mating-type-specific regions, the divergence of the A1- and A2-linked homologs should be roughly proportional to the time since recombination stopped in the region. We therefore sequenced and compared PCR fragments amplified from the two mating types of Sl405 or Sl127 cultures (GenBank accession nos. FI855822FI856001). Sequencing of PCR products showed that 12 (4 A1 and 8 A2) loci have more than one copy, and they were excluded from further analysis. Sequences of 61 loci were identical between the A1 and A2 strains, and four loci demonstrated low total divergence (0.24–0.61%) between the two mating types (otintseva and D. Filatov, unpublished results). Thus, these loci might be located in the recombining part of the mating-type-specific chromosomes. Ten of 75 loci that amplified in both mating types demonstrated multiple polymorphisms fixed between the mating types rather than between the locations. Given that the strains that we used in the analysis originated from two geographically distant locations, it is highly unlikely that multiple polymorphisms distinguishing the A1 and A2 sequences arose purely by chance; thus, these loci are probably located in the nonrecombining mating-type-specific region of the M. violaceum A1 and A2 chromosomes.

TABLE 1

Loci from mating-type-specific chromosomes of M. violaceum used for PCR analysis and genetic map construction
With nonzero A1/A2 divergenceb
LociMating type specific<1%>1%With zero A1/A2 divergencebTotal
A1a52 (1)3 (3)35 (3)45 (7)
A2a62 (0)7 (7)26 (3)41 (10)
Subtotal4 (1)10 (10)
Total1114 (11)61 (6)86 (17)
Open in a separate windowaA1, loci originated from the A1 sex chromosome; A2, loci originated from the A2 sex chromosome.bThe number of loci used for genetic map construction is in parentheses.To confirm the mating-type-specific or pseudoautosomal locations of the loci with and without A1/A2 divergence, we conducted genetic mapping in a family of 99 individuals, 50 of which were of mating type A1 and 49 of mating type A2. The family was generated by a cross between A1 and A2 M. violaceum strains from S. latifolia accessions Sl405 (Sweden) and Sl127 (France), respectively. The choice of strains from geographically distant locations was motivated by the hope of maximizing the number of DNA sequence differences between them that can be used as molecular genetic markers in segregation analysis. We inoculated S. latifolia seedlings with sporidial cultures of both mating types. For inoculation, petri dishes with 12-day-old seedlings of S. latifolia were flooded with 2.5 ml of inoculum suspension. Inoculum suspension consisted of equal volumes of the A1 and A2 sporidial cultures that were mixed and conjugated overnight at 14° under rotation (Biere and Honders 1996; Van Putten et al. 2003). Seedlings were potted 3 days after inoculation. Two months later, teliospores were collected from the flowers of the infected plant and grown in petri dishes on 3.6% potato dextrose agar medium. Haploid sporidia formed after meiosis were isolated and grown as separate cultures for DNA extraction. The mating types of single sporidia cultures were identified as described above. The loci analyzed in the segregation analysis were sequenced in the two parental haploid strains and in 99 (50 A1 and 49 A2) haploid strains that were generated in the cross. Single nucleotide differences between the parental strains were used as molecular genetic markers for segregation analysis in the progeny. The genetic map was constructed using MAPMAKER/EXP v3.0 (Lincoln et al. 1992) and MapDisto v1.7 (http://mapdisto.free.fr/).The resulting genetic map is shown in Figure 2. As expected, no recombination was observed between the 10 loci with diverged A1- and A2-linked copies. In addition, one marker with no A1/A2 divergence, A2-397, was also completely linked to the loci with significant A1/A2 divergence. This locus either may be very tightly linked to the nonrecombining mating-type-specific region or may have been added to that region more recently than the loci that had already accumulated some divergence between the alleles in the two mating types. The mating-type-specific pheromone receptor locus (Devier et al. 2009) and 11 mating-type-specific loci are also located in this nonrecombining region (Figure 2). Interestingly, the cluster of nonrecombining markers is flanked on both sides with markers that recombine in meiosis, demonstrating that there are pseudoautosomal regions on both ends of the mating-type-specific chromosomes.Open in a separate windowFigure 2.—Genetic map of the mating-type-determining chromosome in M. violaceum. Genetic distance (in centimorgans) and the relative positions of the markers are shown to the left and the right of the chromosome, respectively. The position of the nonrecombining region corresponds to the cluster of linked markers shown on the right of the figure. Total A1/A2 divergence is shown in parentheses. Eleven mating-type-specific markers (for which sequences are available from only one mating type), located in the nonrecombining mating-type-specific region, are not shown.Our results demonstrate that although the loci reported by Hood et al. (2004) were isolated from the A1 and A2 chromosomes, most of these loci are not located in the nonrecombining mating-type-specific regions. In fact, the nonrecombining region might be relatively small: of 86 tested fragments, only 21 appeared to be either mating type specific or linked to the mating-type locus. Assuming that these loci represent a random set of DNA fragments isolated from the A1 and A2 chromosomes, it is possible to estimate the size of the nonrecombining region using the binomial distribution: the nonrecombining region is expected to be 24.4% (95% CI: 16.7–33.6%) of the chromosome length. As the sizes of the A1 and A2 chromosomes are ∼3.4 and 4.2 Mb long (Hood 2002; Hood et al. 2004), the nonrecombining region might be ∼1 Mb long.Interestingly, total A1/A2 divergence for the 11 loci with A1- and A2-linked copies mapped to the nonrecombining region varied from 0% to 8.6% (Figure 2). In addition, 11 loci amplified from only one mating type. These genes could represent degenerated genes, some of which degenerated in A1 strains, and some in A2 strains. Alternatively, they might be highly diverged genes, such that the PCR primers amplify only one allele, and not the other. Variation in divergence may be the result of the stepwise cessation of recombination between the A1 and A2 chromosomes in M. violaceum, resembling the evolutionary strata reported for human, chicken, and white campion sex chromosomes (Lahn and Page 1999; Handley et al. 2004; Bergero et al. 2007). However, only the differences between the most and the least diverged loci are statistically significant (Devier et al. 2009), the M. violaceum mating-type region has at least three strata: one oldest stratum, including the pheromone receptor locus; a younger stratum with ∼5–9% A1/A2 divergence; and the youngest stratum with 1–4% divergence between the two mating types. There may also be an additional very recently evolved stratum containing the locus named A2-397, which is also present in all A1 strains tested, with no fixed differences between the A1 and A2 strains (
No. of sites analyzedWithin A1
Within A2
Fixed differences between A1 and A2A1/A2 divergence (%)
LociaSb totalSπ (%)cSπ (%)c
A1/A2 divergence <1%A1-23645630020.4410.44
A1-0456544000040.61
A2-568413220.4820.4800.24
A2-411480210.210010.31
A1/A2 divergence >1%A1-2176679000091.35
A1-12856990010.1881.49
A1-199618130010.16122.02
A2-4223449000092.62
A2-516470140000142.98
A2-404508200030.59173.64
A2-4355062220.3920.39183.95
A2-4734572310.2210.22214.81
A2-4573031710.3300165.54
A2-5755034750.9930.59398.55
Open in a separate windowaA1, loci originated from the A1 sex chromosome; A2, loci originated from the A2 sex chromosome.bS, number of polymorphic sites.cπ (%), average number of differences per 100 nucleotides.

TABLE 3

P-values for the 2 × 2 G-tests for significance of differences in A1/A2 divergence between the loci in the nonrecombining region
LaSbLocusA2-397A1-217A1-128A1-199A2-422A2-516A2-404A2-435A2-473A2-457
5190A2-397
6679A1-2170.006
5698A1-1280.0060.93
61812A1-1990.00070.410.48
3449A2-4220.00030.170.210.51
47014A2-5160.000030.060.0860.280.76
50817A2-40400.0250.0380.150.550.75
50618A2-43500.0150.0240.1040.450.620.86
45721A2-47300.0010.0030.01630.150.210.340.43
30316A2-45700.00090.00170.00970.090.130.2030.260.69
50339A2-5750000.000010.0020.0020.0030.0060.0550.199
Open in a separate windowP-values <0.05 are in boldface type.aL, the length of the region compared.bS, the number of nucleotide differences observed.As most of the loci isolated from the A1 and A2 chromosomes recombine in meiosis, they are not expected to degenerate. Thus, the observation of a higher proportion of TEs in these loci, compared to other chromosomes (Hood et al. 2004), is unlikely to reflect genetic degeneration attributable to a lack of recombination in these loci. A higher abundance of TEs in the sequences isolated from the A1 and A2 chromosomes, as reported by Hood et al. (2004), may simply reflect variation in the TE density across the genome. Thus, it remains to be seen whether M. violaceum mating-type-specific regions degenerate, similar to vertebrate Y (or W) chromosomes, or remain largely intact, as in C. neoformans (Lengeler et al. 2002). If the latter were the case, it may suggest that nonrecombining regions in fungi do not necessarily follow the same degenerative path as animal Y and W chromosomes. The analysis of sequences from the M. violaceum genome (and perhaps other fungal genomes) will hopefully provide the answer to this question.The lack of degeneration of mating-type-specific regions in C. neoformans may be due to the relatively small size of the nonrecombining regions. The 20 genes present in this region may not be sufficient for the operation of such detrimental population genetic processes as background selection or Muller''s ratchet because the speed of these processes depends critically on the number of active genes linked together (Charlesworth 2008). Larger mating-type-specific regions in M. violaceum might contain more genes; thus, more active genetic degeneration may be expected in this species. Indeed, many strains of M. violaceum show haplolethality linked to one of the mating types (Hood and Antonivics 2000; Thomas et al. 2003; Tellier et al. 2005), which may reflect the accumulation of deleterious mutations in the nonrecombining regions around the mating-type loci. Mating-type specificity of the markers that amplified in only A1 or A2 strains in this study may also reflect genetic degeneration.Another factor that may potentially prevent degeneration of genes linked to mating-type loci in fungi is the haploid expression of genes in these regions. In animals, many Y-linked genes have functional homologs on the X chromosome, and loss of the Y-linked gene may be compensated for by expression of the X-linked homologs. The haploid stage in an animal''s life cycle is very short, and very few genes are actively expressed in animal gametes (Schultz et al. 2003). In plants, on the other hand, a significant proportion of the genome is expressed in pollen (da Costa-Nunes and Grossniklaus 2003), and so the loss of Y-linked genes expressed in gametes may be more detrimental than in animals. Indeed, most genes isolated from the white campion X chromosome have intact Y-linked copies (Filatov 2005; Bergero et al. 2007), but due to the small number of genes available, it is still unclear whether genetic degeneration of Y-linked genes is indeed slower in this species (and in plants generally) compared to animal Y chromosomes. Haploid expression could be an even more powerful force in fungi and other organisms with haploid sexes, such as bryophytes, as most genes are expressed in the haploid stage. Further analysis of genetic degeneration in nonrecombining sex- or mating-type-specific regions in fungi and bryophytes will help to shed light on this question.  相似文献   

7.
One- and Two-Locus Population Models With Differential Viability Between Sexes: Parallels Between Haploid Parental Selection and Genomic Imprinting          下载免费PDF全文
Alexey Yanchukov 《Genetics》2009,182(4):1117-1127
A model of genomic imprinting with complete inactivation of the imprinted allele is shown to be formally equivalent to the haploid model of parental selection. When single-locus dynamics are considered, an internal equilibrium is possible only if selection acts in the opposite directions in males and females. I study a two-locus version of the latter model, in which maternal and paternal effects are attributed to the single alleles at two different loci. A necessary condition for the allele frequency equilibria to remain on the linkage equilibrium surface is the multiplicative interaction between maternal and paternal fitness parameters. In this case the equilibrium dynamics are independent at both loci and results from the single-locus model apply. When fitness parameters are additive, analytic treatment was not possible but numerical simulations revealed that stable polymorphism characterized by association between loci is possible only in several special cases in which maternal and paternal fitness contributions are precisely balanced. As in the single-locus case, antagonistic selection in males and females is a necessary condition for the maintenance of polymorphism. I also show that the above two-locus results of the parental selection model are very sensitive to the inclusion of weak directional selection on the individual''s own genotypes.PARENTAL genetic effects refer to the influence of the mother''s and father''s genotypes on the phenotypes of their offspring, not attributable just to the transfer of genes. Examples have been documented across a wide range of areas of organism biology; see, for example, Wade (1998) and and22 in Rasanen and Kruuk (2007). Parental selection is a more formal concept used in theoretical modeling and concerns situations where the fitness of the offspring depends, besides other factors, on the genotypes of its parent(s) (generalizing from Kirkpatrick and Lande 1989).

TABLE 1

Frequencies of genotypes and fitness parameterizations in model 1
Gametes/haploidsFrequency before selectionFitness
ZygoteMaleFemale
(A)AApfpm1 − α1 − δ
(A)a1/2 A 1/2 apf(1 − pm)11
(a)A1/2 a 1/2 A(1 − pf)pm1 − α1 − δ
(a)aA(1 − pf)(1 − pm)11
Open in a separate windowParentheses in the first column indicate maternal genotype (parental selection model) or inactivation of the maternally derived allele (imprinting model). Whether selection occurs at the diploid (first column) or subsequent haploid (second column) stage does not change the resulting allele frequencies.

TABLE 2

Offspring genotypic proportions from different mating types, sorted among four phenotypic groups/combinations of maternal and paternal effects: model 2
Offspring genotypes/phenotypes
Parental genotypes
Paternal (φ = 1)
Joint (φ = 4)
MaleFemaleABAbaBAbABAbaBab
ABAB1
Ab
aB
ab(1−r)/2r/2r/2(1−r)/2
AbAB
Ab1
aBr/2(1−r)/2(1−r)/2r/2
ab
Offspring genotypes/phenotypes
Parental genotypes
Maternal (φ = 2)
None (φ = 3)
MaleFemaleABAbaBAbABAbaBab
aBAB
Abr/2(1 − r)/2(1 − r)/2r/2
aB1
ab
abAB(1 − r)/2(1 − r)/2
Ab
aB
ab1
Open in a separate windowAnother well-known parent-of-origin phenomenon is genomic imprinting. Here, the level of expression of one of the alleles depends on which parent it is inherited from. Often it is difficult to tell apart the phenotypic patterns due to parental effects and genomic imprinting, and thus a problem arises in the process of identifying the candidate genes for such effects (Hager et al. 2008). Analytic methods (Weinberg et al. 1998; Santure and Spencer 2006; Hager et al. 2008) have been developed to quantify subtle differences between the two. In this article, I point out that a simple mathematical model, first suggested for genomic imprinting at a diploid locus, can be interpreted, without any formal changes, to describe parental selection on haploids.While there has been much progress in understanding the evolution of genomic imprinting (Hunter 2007), including advances in modeling (Spencer 2000, 2008), the population genetics theory of parental effects received less attention. Existing major-locus effect models of parental selection are single-locus, two-allele, and mostly concern uniparental (maternal) selection (Wright 1969; Spencer 2003; Gavrilets and Rice 2006; Santure and Spencer 2006), with only one specific case where the fitness effects of both parents interact studied by Gavrilets and Rice (2006). No attempt to extend this theory into multilocus systems has yet been made. Considering a two-locus model with both parents playing a role in selection on the offspring is called for by the observation that many maternal and paternal effects aim at the different traits or different life stages of their progeny. Among birds, for example, body condition soon after hatching is largely determined by the mother, while paternally transmitted sexual display traits develop much later in life (Price 1998). Such effects are therefore unlikely to be regulated within a single locus. Sometimes the effects are on the same trait, but still attributed to different loci: expression of gene Avy that causes the “agouti” phenotype (yellow fur coat and obesity) in mice is enhanced by maternal epigenetic modification (Rakyan et al. 2003), while paternal mutations at the other locus, MommeD4, contribute to a reverse phenotypic pattern in the offspring (Rakyan et al. 2003). The epigenetic state of the murine AxinFu allele is both maternally and paternally inherited (Rakyan et al. 2003).Focusing selection on haploids reduces the number of genotypes that need to be taken into account, while preserving the main properties of the multilocus system. Genes with haploid expression and a potential of parental effects can be found in two major taxonomic kingdoms. A notable candidate is Spam1 in mice, which is expressed during spermogenesis and encodes a factor that enables sperm to penetrate the egg cumulus (Zheng et al. 2001). This gene remains a target for effectively haploid selection, because its product is not shared via cytoplasm bridges between developing spermatides. Mutations at Spam1 alter performance of the male gametes that carry it and might indirectly, perhaps by altering the timing of fertilization, affect the fitness of the zygote. The highest estimated number of mouse genes expressed in the male gametes is currently 2375 (Joseph and Kirkpatrick 2004), and one might expect some of them to have similar paternal effects. Plants go through a profound haploid stage in their life cycles, and genes involved at this stage have an inevitable effect on the fitness of the future generations. In angiosperms, seed development is known to be controlled by both maternal (Chaudhury and Berger 2001; Yadegari and Drews 2004) and paternal (Nowack et al. 2006) effect genes, expressed, respectively, in female and male gametophytes.Under haploid selection, there can be no overdominance, and thus polymorphism is much more difficult to maintain than in diploid selection models (summarized in Feldman 1971). Nevertheless, differential or antagonistic selection between sexes can lead to a new class of stable internal equilibria in the diploid systems (Owen 1953; Bodmer 1965; Mandel 1971; Kidwell et al. 1977; Reed 2007), and I make use of this property in the haploid models developed below. In the experiment by Chippindale and colleagues (Chippindale et al. 2001), ∼75% of the total fitness variation in the adult stage of Drosophila melanogaster was negatively correlated between males and females, which suggests that a substantial portion of the fruit fly expressed genome is under sexually antagonistic selection. I assume that the effect of either parent on the fitness of the individual depends on the sex of the latter, which in respect to modeling is equivalent to the assumption of differential viability between the sexes in the progeny of the same parent(s). Biological systems that satisfy the latter assumptions can be found among colonial green algae: many members of the order Volvocales are haploid except for the short zygotic stage, and during sexual reproduction, they are also dioecious and anisogametic. I return to this example in the discussion. The possibility that genes expressed in animal gametes may be under antagonistic selection between sexes has been discussed (Bernasconi et al. 2004). For example, a (hypothetical) mutation increasing the ATP production in mitochondria would be beneficial in sperm, because of the increased mobility of the latter, but neutral or detrimental in the egg, due to a higher level of oxidative damage to DNA (Zeh and Zeh 2007).My main purpose was to derive conditions for existence and stability of the internal equilibria of the model(s). I begin with a simple one-locus case, which can be analyzed explicitly, and show how these one-locus results can be extended to the case of two recombining loci with multiplicative fitness. Then, I assume an additive relation between the maternal and paternal effect parameters and study the special cases where parental effects are symmetric.  相似文献   

8.
Healing of Euchromatic Chromosome Breaks by Efficient de novo Telomere Addition in Drosophila melanogaster     
Simon W. A. Titen  Kent G. Golic 《Genetics》2010,184(1):309-312
Many arthropod species are infected with maternally inherited endosymbionts that induce a shift in the sex ratio of their hosts by feminizing or killing males (cytoplasmic sex-ratio distorters, or SRDs). These endosymbionts can have profound impacts on evolutionary processes of their hosts. Here, I derive analytical expressions for the coalescent effective size Ne of populations that are infected with SRDs. Irrespective of the type of SRD, Ne for mitochondrial genes is given by the number of infected females. For nuclear genes, the effective population size generally decreases with increasing prevalence of the SRD and can be considerably lower than the actual size of the population. For example, with male-killing bacteria that have near perfect maternal transmission, Ne is reduced by a factor that is given to a good approximation by the proportion of uninfected individuals in the population. The formulae derived here also yield the effective size of populations infected with mutualistic endosymbionts or maternally inherited bacteria that induce cytoplasmic incompatibility, although in these cases, the reduction in Ne is expected to be less severe than for cytoplasmic SRDs.SIMPLE null models are essential in science. In population genetics, this role is filled by the Wright–Fisher model and its retrospective counterpart, the Kingman coalescent. Both of these models have proven to be immensely useful in spite of the fact that natural populations usually violate the assumptions made in these models. The reason for this is that often, the Wright–Fisher model can be rescaled so that it behaves in many important respects like a more complex population model. This rescaling is achieved through the concept of the effective population size, Ne. Roughly speaking, a complex population model is said to have a certain Ne if the haploid Wright–Fisher model with population size Ne experiences the same amount of random genetic drift as the complex model. Reflecting the different ways in which drift can be measured, Ne can be defined in different ways, e.g., as the inbreeding, the variance, or the coalescent effective population size. Different definitions often produce the same value for Ne, but may also yield drastically different numbers (Kimura and Crow 1963).The coalescent effective population size is defined through the factor by which time needs to be rescaled in a complex population model to produce the standard coalescent with time scale given by the population size N (Nordborg and Krone 2002). It has been argued that this is the most useful definition for Ne because “the coalescent essentially embodies all of the information that can be found in sampled genetic data” (Sjödin et al. 2005). More recently, Wakeley and Sargsyan (2009) have proposed two extensions of the coalescent effective population size in which they advocate including a mutation parameter in the definition and also allowing for a nonlinear relationship between Ne and N.One frequently encountered feature in natural populations that complicates population genetics is infection with maternally inherited endosymbionts. In particular, many arthropod species harbor a great number of phylogenetically diverse microorganisms that influence their hosts'' biology in different ways (Bourtzis and Miller 2003; Bourtzis and Miller 2006; Bourtzis and Miller 2009). Because of their maternal transmission, many of these microorganisms—for example, the bacteria Wolbachia pipientis and Cardinium hertigii—have evolved intricate manipulations of their hosts'' reproductive system that allows them to spread in a host population through exploitation of male hosts (reproductive parasitism, reviewed in Engelstädter and Hurst 2009a). Most manipulations involve a shift in the sex ratio of their hosts (both primary and at the population level), and the inducing endosymbionts are consequently referred to as cytoplasmic sex-ratio distorters (SRDs). In some species, genetic males develop into females if they are infected (“feminization”: Martin et al. 1973; Rigaud 1997; Bouchon et al. 1998; Hiroki et al. 2002). More commonly, infected males are killed by the endosymbionts early in their development (male killing: reviewed in Hurst et al. 2003). The adaptive advantage of this strategy is seen in an early fitness boost in the surviving females in a brood, for example, through reduced sibling competition or cannibalism of the dead brothers (Hurst 1991; Hurst and Majerus 1993; Jaenike et al. 2003). Some examples for species infected with male-killing or feminizing SRDs are given in Huigens and Stouthamer 2003).

TABLE 1

Empirical examples for cytoplasmic SRDs with parameter estimates
HostSRDReferences
Acrea encedona (butterfly)Wolbachia≈ 0≈ 0.5≈ 0Jiggins et al. (2002)
Adalia bipunctatab (Ladybird beetle)Rickettsia0.0760.5060.076Hurst et al. (1993)
Drosophila innubilabWolbachia0.0130.5090.037Dyer and Jaenike (2004)
Gammarus duebenic (freshwater shrimp)microsporidium0.1270.7060.167Dunn et al. (1993)
Hypolimnas bolinaa (butterfly)Wolbachia≈ 0≈ 0.5≈ 0Dyson et al. (2002)
Open in a separate windowWith the exception of the feminizing microsporidia in G. duebeni, all SRDs are male-killing bacteria.aIn these two butterfly species, maternal transmission and male-killing penetrance is very close to perfect, so that virtually no sons or uninfected daughters are produced by infected mothers. The fact that the infection has not spread to fixation in these species suggests that the fitness benefit of surviving siblings in a brood is absent or very low, leading to the prediction that infected females produce about as many daughters as uninfected females ().bIn these two species, parameter estimates were obtained from the transmission rate and the prevalence () reported in the respective references (see Equation 2). In the case of D. innubila, I used the data from the 2002 sampling in Dyer and Jaenike (2004), as this was the largest sample and lay in between the other two samples with regard to infection prevalence.cTo calculate the parameter values given for this species in Dunn et al. (1993, Dunn et al. 1993) and discarded occasionally produced intersexes. Also note that the primary sex ratio in uninfected G. duebeni is determined by environmental cues and can therefore deviate from 1:1.Previous theoretical studies indicate that cytoplasmic SRDs will have a strong impact on evolutionary processes for both mitochondrial and nuclear host genes. This is essentially because the host population consists of different classes of individuals (male/female, infected/uninfected) with different reproductive success. Johnstone and Hurst (1996) showed that genetic variation in mtDNA is expected to be strongly reduced during the spread of male-killing bacteria. After the male killers have reached a stable equilibrium in the population, mtDNA variation will recover, but will still be permanently reduced to a value that approximately corresponds to the expected variation if the population consisted only of infected females. In other words, the equilibrium Ne equals approximately the number of infected females in this case. Conversely, for nuclear host genes, Engelstädter and Hurst (2007) showed through computer simulations that to a good approximation, a male-killer infected population behaves as if only uninfected individuals were present.Here, I derive analytical expressions for the coalescent effective size of host populations infected with cytoplasmic SRDs at equilibrium frequency. This is done for both mitochondrial and nuclear genes. The approach is considerably more general than in the two above-mentioned previous studies in that not only male killing, but also feminizing and even endosymbionts without sex-ratio distorting activity, are covered. However, diploid hosts are assumed throughout this article, so that the derivation for Ne in populations infected with parthenogenesis-inducing bacteria is left for future investigations.  相似文献   

9.
Genetic Testing of the Hypothesis That Hybrid Male Lethality Results From a Failure in Dosage Compensation     
Daniel A. Barbash 《Genetics》2010,184(1):313-316
  相似文献   

10.
Variation in Genomic Recombination Rates Among Heterogeneous Stock Mice          下载免费PDF全文
Beth L. Dumont  Karl W. Broman  Bret A. Payseur 《Genetics》2009,182(4):1345-1349
We used a large panel of pedigreed, genetically admixed house mice to study patterns of recombination rate variation in a leading mammalian model system. We found considerable inter-individual differences in genomic recombination rates and documented a significant heritable component to this variation. These findings point to clear variation in recombination rate among common laboratory strains, a result that carries important implications for genetic analysis in the house mouse.THE rate of recombination—the amount of crossing over per unit DNA—is a key parameter governing the fidelity of meiosis. Recombination rates that are too high or too low frequently give rise to aneuploid gametes or prematurely arrest the meiotic cell cycle (Hassold and Hunt 2001). As a consequence, recombination rates should experience strong selective pressures to lie within the range defined by the demands of meiosis (Coop and Przeworski 2007). Nonetheless, classical genetic studies in Drosophila (Chinnici 1971; Kidwell 1972; Brooks and Marks 1986), crickets (Shaw 1972), flour beetles (Dewees 1975), and lima beans (Allard 1963) have shown that considerable inter-individual variation for recombination rate is present within populations. Recent studies examining the transmission of haplotypes in human pedigrees have corroborated these findings (Broman et al. 1998; Kong et al. 2002; Coop et al. 2008).Here, we use a large panel of heterogeneous stock (HS) mice to study variation in genomic recombination rates in a genetic model system. These mice are genetically admixed, derived from an initial generation of pseudorandom mating among eight common inbred laboratory strains (DBA/2J, C3H/HeJ, AKR/J, A/J, BALB/cJ, CBA/J, C57BL/6J, and LP/J), followed by >50 generations of pseudorandom mating in subsequent hybrid cohorts (Mott et al. 2000; Demarest et al. 2001). The familial relationships among animals in recent generations were tracked to organize the mice into pedigrees. In total, this HS panel includes ∼2300 animals comprising 85 families, 8 of which span multiple generations. The remainder consists of nuclear families (sibships) that range from 1 to 34 sibs, with an average of 9.6 sibs (Valdar et al. 2006) (Mott et al. 2000; Demarest et al. 2001; Shifman et al. 2006).

TABLE 1

Heterogeneous stock mouse pedigrees
PedigreePedigree classNo. of nonoverlapping sibships in the pedigreeNo. of retained sibshipsNo. of meioses
1Multigenerational1717464
2Multigenerational2720728
3Multigenerational2319602
4Multigenerational149254
5Multigenerational119242
6Multigenerational5368
7Multigenerational43100
8Multigenerational2116
9Sibshipa2120
32–85Sibship511146
Total1801323640
Open in a separate windowaThis family was composed of two sibships sharing a common mother but with different fathers.With the exception of several founding individuals, most of these HS mice have been genotyped at 13,367 single nucleotide polymorphisms (SNPs) across the genome (available at http://gscan.well.ox.ac.uk/). Although the publicly available HS genotypes have passed data quality filters (Shifman et al. 2006), we took several additional measures to ensure the highest possible accuracy of base calls. First, data were cleansed of all non-Mendelian inheritances, and genotypes with quality scores <0.4 were removed. Genotypes that resulted in tight (<10 cM in sex-specific distance) double recombinants were also omitted because strong positive crossover interference in the mouse renders such closely spaced crossovers biologically very unlikely (Broman et al. 2002). A total of 10,195 SNPs (including 298 on the X chromosome) passed these additional quality control criteria; the results presented below consider only this subset of highly accurate (>99.98%) and complete (<0.01% missing) genotypes. The cleaned data are publicly available (at http://cgd.jax.org/mousemapconverter/).We used the chrompic program within CRI-MAP (Lander and Green 1987; Green et al. 1990) to estimate the number of recombination events in parental meioses. The algorithm implemented in chrompic first phases parent and offspring genotypes using a maximum-likelihood approach. Next, recombination events occurring in the parental germline are identified by comparing parent and offspring haplotypes across the genome (Green et al. 1990). For example, a haplotype that first copies from one maternal chromosome and then switches to copying from the other maternal chromosome signals a recombination event in the maternal germline.chrompic is very memory intensive and cannot handle the multigenerational pedigrees and the large sibships included in the HS panel. To circumvent these computational limitations, several modifications to the data were implemented. First, the eight multigenerational pedigrees were split into 102 nonoverlapping sibships, retaining grandparental information when available (Cox et al. 2009). Finally, large sibships were subdivided: sibships with >13 progeny were split into two groups: those with >26 progeny were split into three groups and those with >39 sibs were split into four groups. Partitioning large sibships by units of 10, 11, or 12, rather than 13, had no effect on the estimation of crossover counts, suggesting that the estimates were robust to the unit of subdivision. These subdivided families were used only for haplotype inference; all other analyses treated whole sibships as focal units. In total, we analyzed 132 nonoverlapping sibships, ranging in size from 2 to 48 sibs (mean = 13.9). This data set encompassed 3640 meioses—300–2000% more meioses than previously studied human pedigrees (Broman et al. 1998; Kong et al. 2002; Coop et al. 2008)—providing excellent power to detect recombination rate variation among individuals.The recombination rate for the maternal (or paternal) parent of a given sibship was estimated as the average number of recombination events in the haploid maternal (or paternal) genomes transmitted to her (or his) offspring. Our analyses treat males and females separately, as previous observations in mice (Murray and Snell 1945; Mallyon 1951; Reeves et al. 1990; Dietrich et al. 1996; Shifman et al. 2006; Paigen et al. 2008), along with findings from this study, point to systematically higher recombination rates in female than in male mice (this study: P < 2.2 × 10−16, Mann–Whitney U-Test comparing autosomal crossover counts in the 131 HS females to those in the 131 HS males).There is considerable recombination rate heterogeneity among the 131 mothers and 131 fathers in the HS pedigrees (Figure 1). The female with the highest recombination rate had an average of nearly twice as many crossovers per meiosis compared with the lowest (female range: 9.0–17.3; mean = 13.3; SD = 3.28). Similarly, the least actively recombining male had only 55% the amount of recombination as the male with the highest recombination rate (male range: 7.7–14.7; mean = 11.7; SD = 2.76). These average values are similar to previously reported recombination counts in house mice, determined using both cytological (Dumas and Britton-Davidian 2002; Koehler et al. 2002) and genetic (Dietrich et al. 1996) approaches. Note that the recombination rates that we report reflect the number of exchange events visible in genetic data. Under the assumption of no chromatid interference, the expected number of crossovers that occur at meiosis is equal to twice these values.Open in a separate windowFigure 1.—Variation in recombination frequency in HS mice. The mean number of recombination events per transmitted gamete in each mother (A; n = 131) and father (B; n = 131) was inferred by comparing parent and offspring genotypes at >10,000 autosomal and X-linked markers using the CRIMAP chrompic computer program. Error bars span ±2 SEs.To test for variation in recombination within the HS females and within the HS males, we performed a one-way ANOVA using parental identity as the factor and the recombination count for a single haploid genome transmission on the pedigree as the response variable. Significance of the resultant F-statistic was empirically assessed by randomizing parental identity with respect to individual recombination counts, recomputing the F-statistic on the permuted data set, and determining the quantile position of the observed F-statistic along the distribution of 106 F-statistics derived from randomization. There is highly significant variation for genomic recombination rate among HS females (F = 1.7842, P < 10−6; Figure 1A) and males (F = 2.3103, P < 10−6; Figure 1B).We next examined patterns of recombination rate inheritance using the eight complex families to test for heritability of this trait. We fit a polygenic model of inheritance using the polygenic command within SOLAR v.4, accounting for the uneven relatedness among individuals through a matrix of pairwise coefficients of relatedness (Almasy and Blangero 1998). Sex was included as a covariate in the model to account for the well-established differences between male and female recombination rates in mice (Murray and Snell 1945; Mallyon 1951; Reeves et al. 1990; Dietrich et al. 1996; Shifman et al. 2006; Paigen et al. 2008). Recombination rates show significant narrow-sense heritability (h2 = 0.46; SE = 0.20; P = 0.008), indicating that variation for recombination rate among HS mice is partly attributable to additive genetic variation. This result agrees with previous evidence for genetic effects on recombination rate variation in the house mouse (Reeves et al. 1990; Shiroishi et al. 1991; Koehler et al. 2002).In summary, we have shown that HS mice differ significantly in their genomic recombination rates and have demonstrated that this variation is heritable. These findings indicate that interstrain variation for genomic average recombination rate exists among at least two of the eight progenitor strains of the HS stock, mirroring observations of significant variation among inbred laboratory strains for many other quantitative characters (Grubb et al. 2009). Indeed, cytological analyses have already revealed significant differences in recombination frequencies between A/J and C57BL/6J males (Koehler et al. 2002), two of the HS founding strains.This interstrain variation in genomic recombination rate carries important practical implications for genetic analysis in the house mouse. Most notably, crosses using inbred mouse strains with high recombination rates will provide higher mapping resolution than crosses using strains with reduced recombination rates. However, the strategic use of high-recombination-rate strains will not necessarily expedite the fine mapping of loci. The distribution of recombination events in mice is not uniform across chromosomes and appears to be strain specific (Paigen et al. 2008; Grey et al. 2009; Parvanov et al. 2009).The history of the classical inbred mouse strains as inferred from pedigrees (Beck et al. 2000), sequence comparisons to wild mice (Salcedo et al. 2007), and genomewide phylogenetic analyses (Frazer et al. 2007; Yang et al. 2007) suggests that much of the interstrain variation for recombination rate arises from genetic polymorphism among Mus domesticus individuals in nature. However, many other factors have likely shaped recombination rate variation among the classical strains, including inbreeding, artificial selection, and hybridization with closely related species (Wade and Daly 2005). These aspects of the laboratory mouse''s history challenge comparisons between recombination rate variation in the HS panel and human populations and provide strong motivation for studies of recombination rate variation in natural populations of house mice.Although we find a strong genetic component to inter-individual variation in recombination rate, a large fraction (∼54%) of the phenotypic variation for recombination is not explained by additive genetic variation alone. Sampling error and other forms of genetic variation (e.g., dominance and epistasis) likely combine to account for some of the residual variation. In addition, micro-environmental differences within the laboratory setting (Koren et al. 2002) and life history differences among families, including parental age (Koehler et al. 2002; Kong et al. 2004), might contribute to variation in recombination rates among the HS mice.Identifying the genetic loci that underlie recombination rate differences among the HS mice (and hence in the eight founding inbred strains) presents a logical next step in the research program initiated here. The complicated pedigree structure, relatively small number of animals with recombination rate estimates (n = 262), and potentially sex-specific genetic architecture of this trait (Kong et al. 2008; Paigen et al. 2008) will pose challenges to this analysis. Nonetheless, dissecting the genetic basis of recombination rate variation is a pursuit motivated by its potential to lend key insights into several enduring questions. Why do males and females differ in the rate and distribution of crossover events? What are the evolutionary mechanisms that give rise to intraspecific polymorphism and interspecific divergence for recombination rate? What are the functional consequences of recombination rate variation? Alternative experimental approaches, including those that combine the power of QTL mapping with immunocytological assays for measuring recombination rates in situ (Anderson et al. 1999), promise to offer additional clues onto the genetic mechanisms that give rise to variation in this important trait.  相似文献   

11.
Interaction Between Eye Pigment Genes and Tau-Induced Neurodegeneration in Drosophila melanogaster     
Surendra S. Ambegaokar  George R. Jackson 《Genetics》2010,186(1):435-442
  相似文献   

12.
A Two-Pathway Analysis of Meiotic Crossing Over and Gene Conversion in Saccharomyces cerevisiae     
Franklin W. Stahl  Henriette M. Foss 《Genetics》2010,186(2):515-536
Several apparently paradoxical observations regarding meiotic crossing over and gene conversion are readily resolved in a framework that recognizes the existence of two recombination pathways that differ in mismatch repair, structures of intermediates, crossover interference, and the generation of noncrossovers. One manifestation of these differences is that simultaneous gene conversion on both sides of a recombination-initiating DNA double-strand break (“two-sidedness”) characterizes only one of the two pathways and is promoted by mismatch repair. Data from previous work are analyzed quantitatively within this framework, and a molecular model for meiotic double-strand break repair based on the concept of sliding D-loops is offered as an efficient scheme for visualizing the salient results from studies of crossing over and gene conversion, the molecular structures of recombination intermediates, and the biochemical competencies of the proteins involved.EUKARYOTES transit from the diplophase to the haplophase via meiosis, which is associated with a number of interrelated processes, including crossing over and gene conversion. These processes involve meiosis-specific, programmed DNA double-strand breaks (DSBs) and their repair (DSBr). DSBr, in turn, is associated with mismatched base pairs and their rectification, referred to as “mismatch repair” or MMR (Bishop et al. 1987). Current efforts to accommodate both the genetic and molecular phenomena associated with meiotic DSBr in yeast (Saccharomyces cerevisiae) have been thoroughly reviewed (e.g., Hollingsworth and Brill 2004; Hoffmann and Borts 2004; Surtees et al. 2004; Hunter 2007; Berchowitz and Copenhaver 2010), but none of the reviews commits to an overall picture with quantitative predictions. This work aims to remedy that lack. Specifically, we have made use of salient published studies to develop, step-by-step, a comprehensive model of meiotic DSBr and MMR. The main features of this model are summarized in FeaturesPairing pathwayDisjunction pathwayProductsCrossovers and noncrossoversCrossovers onlyCrossover InterferenceNo positive interferencePositive interferenceMsh4–Msh5 dependenceNoneTotalBimolecular intermediateLong with junctions not fully ligatedShort with fully ligated Holliday junctionsInvasion heteroduplexPartly ephemeralEphemeralMMR at invasion and annealingDependent on Msh2 and Mlh1NoneMMR near the DSB siteDirected by 3′ invading and annealing endsMlh1 dependent; directed by junction resolutionRole of Msh2 in MMRRecognizes mismatches and attracts Mlh1NoneRole of Msh4–Msh5 in MMRNoneAttracts Mlh1Open in a separate window  相似文献   

13.
Mouse Models of Osteoarthritis: A Summary of Models and Outcomes Assessment     
Sabine Drevet  Bertrand Favier  Emmanuel Brun  Gaëtan Gavazzi  Bernard Lardy 《Comparative medicine》2022,72(1):3
Osteoarthritis (OA) is a multidimensional health problem and a common chronic disease. It has a substantial impact on patient quality of life and is a common cause of pain and mobility issues in older adults. The functional limitations, lack of curative treatments, and cost to society all demonstrate the need for translational and clinical research. The use of OA models in mice is important for achieving a better understanding of the disease. Models with clinical relevance are needed to achieve 2 main goals: to assess the impact of the OA disease (pain and function) and to study the efficacy of potential treatments. However, few OA models include practical strategies for functional assessment of the mice. OA signs in mice incorporate complex interrelations between pain and dysfunction. The current review provides a comprehensive compilation of mouse models of OA and animal evaluations that include static and dynamic clinical assessment of the mice, merging evaluation of pain and function by using automatic and noninvasive techniques. These new techniques allow simultaneous recording of spontaneous activity from thousands of home cages and also monitor environment conditions. Technologies such as videography and computational approaches can also be used to improve pain assessment in rodents but these new tools must first be validated experimentally. An example of a new tool is the digital ventilated cage, which is an automated home-cage monitor that records spontaneous activity in the cages.

Osteoarthritis (OA) is a multidimensional health problem and a common chronic disease.36 Functional limitations, the absence of curative treatments, and the considerable cost to society result in a substantial impact on quality of life.76 Historically, OA has been described as whole joint and whole peri-articular diseases and as a systemic comorbidity.9,111 OA consists of a disruption of articular joint cartilage homeostasis leading to a catabolic pathway characterized by chondrocyte degeneration and destruction of the extracellular matrix (ECM). Low-grade chronic systemic inflammation is also actively involved in the process.42,92 In clinical practice, mechanical pain, often accompanied by a functional decline, is the main reason for consultations. Recommendations to patients provide guidance for OA management.22, 33,49,86 Evidence-based consensus has led to a variety of pharmacologic and nonpharmacologic modalities that are intended to guide health care providers in managing symptomatic patients. Animal-based research is of tremendous importance for the study of early diagnosis and treatment, which are crucial to prevent the disease progression and provide better care to patients.The purpose of animal-based OA research is 2-fold: to assess the impact of the OA disease (pain and function) and to study the efficacy of a potential treatment.18,67 OA model species include large animals such as the horse, goat, sheep, and dog, whose size and anatomy are expected to better reflect human joint conditions. However, small animals such as guinea pig, rabbit, mouse, and rat represent 77% of the species used.1,87 In recent years, mice have become the most commonly used model for studying OA. Mice have several advantageous characteristics: a short development and life span, easy and low-cost breeding and maintenance, easy handling, small joints that allow histologic analysis of the whole joint,32 and the availability of genetically modified lines.108 Standardized housing, genetically defined strains and SPF animals reduce the genetic and interindividual acquired variability. Mice are considered the best vertebrate model in terms of monitoring and controlling environmental conditions.7,14,15,87 Mouse skeletal maturation is reached at 10 wk, which theoretically constitutes the minimal age at which mice should be entered into an OA study.64,87,102 However, many studies violate this limit by testing mice at 8 wk of age.Available models for OA include the following (32,111 physical activity and exercise induced OA; noninvasive mechanical loading (repetitive mild loading and single-impact injury); and surgically induced (meniscectomy models or anterior cruciate ligament transection). The specific model used would be based on the goal of the study.7 For example, OA pathophysiology, OA progression, and OA therapies studies could use spontaneous, genetic, surgical, or noninvasive models. In addition, pain studies could use chemical models. Lastly, post-traumatic studies would use surgical or noninvasive models; the most frequently used method is currently destabilization of the medial meniscus,32 which involves transection of the medial meniscotibial ligament, thereby destabilizing the joint and causing instability-driven OA. An important caveat for mouse models is that the mouse and human knee differ in terms of joint size, joint biomechanics, and histologic characteristics (layers, cellularity),32,64 and joint differences could confound clinical translation.10 Table 1. Mouse models of osteoarthritis.
ModelsProsCons
SpontaneousWild type mice7,9,59,67,68,70,72,74,80,85,87,115,118,119,120- Model of aging phenotype
- The less invasive model
- Physiological relevance: mimics human pathogenesis
- No need for technical expertise
- No need for specific equipment
- Variability in incidence
- Large number of animals at baseline
- Long-term study: Time consuming (time of onset: 4 -15 mo)
- Expensive (husbandry)
Genetically modified mice2,7,25,40,50,52,67,72,79,80, 89,120- High incidence
- Earlier time of onset: 18 wk
- No need for specific equipment
- Combination with other models
- Time consuming for the strain development
- Expensive
Chemical- inducedMono-iodoacetate injection7,11,46,47,60,66,90,91,101,128- Model of pain-like phenotype
- To study mechanism of pain and antalgic drugs
- Short-term study: Rapid progression (2-7 wk)
- Reproducible
- Low cost
- Need for technical expertise
- Need for specific equipment
- Systemic injection is lethal
- Destructive effect: does not allow to study the early phase of pathogenesis
Papain injection66,67,120- Short-term study: rapid progression
- Low cost
- Need for technical expertise
- Need for specific equipment
- Does not mimic natural pathogenesis
Collagenase injection7,65,67,98- Short-term study: rapid progression (3 wk)
- Low cost
- Need for technical expertise
- Need for specific equipment
- Does not mimic natural pathogenesis
Non-invasiveHigh-fat diet (Alimentary induced obesity model)5,8,43,45,57,96,124Model of metabolic phenotype
No need for technical expertise
No need for specific equipment
Reproducible
Long-term study: Time consuming (8 wk–9 mo delay)
Expensive
Physical activity and exercise model45,73Model of post traumatic phenotype
No need for technical expertise
Long-term study: time consuming (18 mo delay)
Expensive
Disparity of results
Mechanical loading models Repetitive mild loading models Single-impact injury model7,16,23,24, 32,35,104,105,106Model of post traumatic phenotype
Allow to study OA development
Time of onset: 8-10 wk post injury
Noninvasive
Need for technical expertise
Need for specific equipment
Heterogeneity in protocol practices
Repetitive anesthesia required or ethical issues
SurgicalOvariectomy114Contested.
Meniscectomy model7,32,63,67,87 Model of post traumatic phenotype
High incidence
Short-term study: early time of onset (4 wk from surgery)
To study therapies
Need for technical expertise
Need for specific equipment
Surgical risks
Rapid progression compared to human
Anterior cruciate ligament transection (ACLT)7,39,40,61,48,67,70,87,126Model of posttraumatic phenotype
High incidence
Short-term study: early time of onset (3-10 wk from surgery)
Reproducible
To study therapies
Need for technical expertise
Need for specific equipment
Surgical risks
Rapid progression compared to human
Destabilization of medial meniscus (DMM)7,32,39,40Model of post traumatic phenotype
High incidence
Short-term study: early time of onset (4 wk from surgery)
To study therapies
The most frequently used method
Need for technical expertise
Need for specific equipment
Surgical risks
Rapid progression compared to human
Open in a separate windowSince all animal models have strengths and weaknesses, it is often best to plan using a number of models and techniques together to combine the results.In humans, the lack of correlation between OA imaging assessment and clinical signs highlights the need to consider the functional data and the quality of life to personalize OA management. Clinical outcomes are needed to achieve 2 main goals: to assess the impact of the OA in terms of pain and function and to study the efficacy of treatments.65 Recent reviews offer few practical approaches to mouse functional assessment and novel approaches to OA models in mice.7,32,67,75,79,83,87, 100,120 This review will focus on static and dynamic clinical assessment of OA using automatic and noninvasive emerging techniques (Test nameTechniquesKind of assessmentOutputSpecific equipment requiredStatic measurementVon Frey filament testingCalibrated nylon filaments of various thickness (and applied force) are pressed against the skin of the plantar surface of the paw in ascending order of forceStimulus- evoked pain-like behavior
Mechanical stimuli - Tactile allodynia
The most commonly used testLatency to paw withdrawal
and
Force exerted are recordedYesKnee extension testApply a knee extension on both the intact and affected knee
or
Passive extension range of the operated knee joint under anesthesiaStimulus-evoked pain-like behaviorNumber of vocalizations evoked in 5 extensionsNoneHotplateMouse placed on hotplate. A cutoff latency has been determined to avoid lesionsStimulus-evoked pain-like behavior
Heat stimuli- thermal sensitivityLatency of paw withdrawalYesRighting abilityMouse placed on its backNeuromuscular screeningLatency to regain its footingNoneCotton swab testBringing a cotton swab into contact with eyelashes, pinna, and whiskersStimulus-evoked pain-like behavior
Neuromuscular screeningWithdrawal or twitching responseNoneSpontaneous activitySpontaneous cage activityOne by one the cages must be laid out in a specific platformSpontaneous pain behavior
Nonstimulus evoked pain
ActivityVibrations evoked by animal movementsYesOpen field analysisExperiment is performed in a clear chamber and mice can freely exploreSpontaneous pain behavior
Nonstimulus evoked pain
Locomotor analysisPaw print assessment
Distance traveled, average walking speed, rest time, rearingYesGait analysisMouse is placed in a specific cage equipped with a fluorescent tube and a glass plate allowing an automated quantitative gait analysisNonstimulus evoked pain
Gait analysis
Indirect nociceptionIntensity of the paw contact area, velocity, stride frequency, length, symmetry, step widthYesDynamic weight bearing systemMouse placed is a specific cage. This method is a computerized capacitance meter (similar to gait analysis)Nonstimulus evoked pain
Weight-bearing deficits
Indirect nociceptionBody weight redistribution to a portion of the paw surfaceYesVoluntary wheel runningMouse placed is a specific cage with free access to stainless steel activity wheels. The wheel is connected to a computer that automatically record dataNonstimulus evoked pain
ActivityDistance traveled in the wheelYesBurrowing analysisMouse placed is a specific cage equipped with steel tubes (32 cm in length and 10 cm in diameter) and quartz sand in Plexiglas cages (600 · 340x200 mm)Nonstimulus evoked pain
ActivityAmount of sand burrowedYesDigital video recordingsMouse placed is a specific cage according to the toolNonstimulus evoked pain
Or
Evoked painScale of pain or specific outcomeYesDigital ventilated cage systemNondisrupting capacitive-based technique: records spontaneous activity 24/7, during both light and dark phases directly from the home cage rackSpontaneous pain behavior
Nonstimulus evoked pain
Activity-behaviorDistance walked, average speed, occupation front, occupation rear, activation density.
Animal locomotion index, animal tracking distance, animal tracking speed, animal running wheel distance and speed or rotationYesChallenged activityRotarod testGradual and continued acceleration of a rotating rod onto which mice are placedMotor coordination
Indirect nociceptionRotarod latency: riding time and speed with a maximum cut off.YesHind limb and fore grip strengthMouse placed over a base plate in front of a connected grasping toolMuscle strength of limbsPeak force, time resistanceYesWire hang analysisSuspension of the mouse on the wire and start the timeMuscle strength of limbs: muscle function and coordinationLatency to fall grippingNone
(self -constructed)
Open in a separate windowPain cannot be directly measured in rodents, so methods have been developed to quantify “pain-like” behaviors. The clinical assessment of mice should be tested both before and after the intervention (induced-OA ± administration of treatment) to take into account the habituation and establish a baseline to compare against.  相似文献   

14.
On the Classification of Epistatic Interactions     
Hong Gao  Julie M. Granka  Marcus W. Feldman 《Genetics》2010,184(3):827-837
Modern genomewide association studies are characterized by the problem of “missing heritability.” Epistasis, or genetic interaction, has been suggested as a possible explanation for the relatively small contribution of single significant associations to the fraction of variance explained. Of particular concern to investigators of genetic interactions is how to best represent and define epistasis. Previous studies have found that the use of different quantitative definitions for genetic interaction can lead to different conclusions when constructing genetic interaction networks and when addressing evolutionary questions. We suggest that instead, multiple representations of epistasis, or epistatic “subtypes,” may be valid within a given system. Selecting among these epistatic subtypes may provide additional insight into the biological and functional relationships among pairs of genes. In this study, we propose maximum-likelihood and model selection methods in a hypothesis-testing framework to choose epistatic subtypes that best represent functional relationships for pairs of genes on the basis of fitness data from both single and double mutants in haploid systems. We gauge the performance of our method with extensive simulations under various interaction scenarios. Our approach performs reasonably well in detecting the most likely epistatic subtype for pairs of genes, as well as in reducing bias when estimating the epistatic parameter (ɛ). We apply our approach to two available data sets from yeast (Saccharomyces cerevisiae) and demonstrate through overlap of our identified epistatic pairs with experimentally verified interactions and functional links that our results are likely of biological significance in understanding interaction mechanisms. We anticipate that our method will improve detection of epistatic interactions and will help to unravel the mysteries of complex biological systems.UNDERSTANDING the nature of genetic interactions is crucial to obtaining a more complete picture of complex biological systems and their evolution. The discovery of genetic interactions has been the goal of many researchers studying a number of model systems, including but not limited to Saccharomyces cerevisiae, Caenorhabditis elegans, and Escherichia coli (You and Yin 2002; Burch et al. 2003; Burch and Chao 2004; Tong et al. 2004; Drees et al. 2005; Sanjuán et al. 2005; Segre et al. 2005; Pan et al. 2006; Zhong and Sternberg 2006; Jasnos and Korona 2007; St. Onge et al. 2007; Decourty et al. 2008). Recently, high-throughput experimental approaches, such as epistatic mini-array profiles (E-MAPs) and genetic interaction analysis technology for E. coli (GIANT-coli), have enabled the study of epistasis on a large scale (Schuldiner et al. 2005, 2006; Collins et al. 2006, 2007; Typas et al. 2008). However, it remains unclear whether the computational and statistical methods currently in use to identify these interactions are indeed the most appropriate.The study of genetic interaction, or “epistasis,” has had a long and somewhat convoluted history. Bateson (1909) first used the term epistasis to describe the ability of a gene at one locus to “mask” the mutational influence of a gene at another locus (Cordell 2002). The term “epistacy” was later coined by Fisher (1918) to denote the statistical deviation of multilocus genotype values from an additive linear model for the value of a phenotype (Phillips 1998, 2008).These origins are the basis for the two main current interpretations of epistasis. The first, as introduced by Bateson (1909), is the “biological,” “physiological,” or “compositional” form of epistasis, concerned with the influence of an individual''s genetic background on an allele''s effect on phenotype (Cheverud and Routman 1995; Phillips 1998, 2008; Cordell 2002; Moore and Williams 2005). The second interpretation, attributed to Fisher, is “statistical” epistasis, which in its linear regression framework places the phenomenon of epistasis in the context of a population (Wagner et al. 1998; Wade et al. 2001; Wilke and Adami 2001; Moore and Williams 2005; Phillips 2008). Each of these approaches is equally valid in studying genetic interactions; however, confusion still exists about how to best reconcile the methods and results of the two (Phillips 1998, 2008; Cordell 2002; Moore and Williams 2005; Liberman and Feldman 2006; Aylor and Zeng 2008).Aside from the distinction between the statistical and the physiological definitions of epistasis, inconsistencies exist when studying solely physiological epistasis. For categorical traits, physiological epistasis is clear as a “masking” effect. When noncategorical or numerical traits are measured, epistasis is defined as the deviation of the phenotype of the multiple mutant from that expected under independence of the underlying genes.The “expectation” of the phenotype under independence, that is, in the absence of epistasis, is not defined consistently between studies. For clarity, consider epistasis between pairs of genes and, without loss of generality, consider fitness as the phenotype. The first commonly used definition of independence, originating from additivity, defines the effect of two independent mutations to be equal to the sum of the individual mutational effects. A second, motivated by the use of fitness as a phenotype, defines the effect of the two mutations as the product of the individual effects (Elena and Lenski 1997; Desai et al. 2007; Phillips 2008). A third definition of independence has been referred to as “minimum,” where alleles at two loci are independent if the double mutant has the same fitness as the less-fit single mutant. Mani et al. (2008) claim that this has been used when identifying pairwise epistasis by searching for synthetic lethal double mutants (Tong et al. 2001, 2004; Pan et al. 2004, 2006; Davierwala et al. 2005). A fourth is the “Log” definition presented by Mani et al. (2008) and Sanjuan and Elena (2006). The less-frequently used “scaled ɛ” (Segre et al. 2005) measure of epistasis takes the multiplicative definition of independence with a scaling factor.These different definitions of independence are partly due to distinct measurement “scales.” For some traits, a multiplicative definition of independence may be necessary to identify epistasis between two genes, whereas for other traits, additivity may be appropriate (Falconer and Mackay 1995; Wade et al. 2001; Mani et al. 2008; Phillips 2008). An interaction found under one independence definition may not necessarily be found under another, leading to different biological conclusions (Mani et al. 2008).Mani et al. (2008) suggest that there may be an “ideal” definition of independence for all gene pairs for identifying functional relationships. However, it is plausible that different representations of independence for two genes may reflect different biological properties of the relationship (Kupper and Hogan 1978; Rothman et al. 1980). “Two categories of general interest [the additive and multiplicative definitions, respectively] are those in which etiologic factors act interchangeably in the same step in a multistep process, or alternatively act at different steps in the process” (Rothman et al. 1980, p. 468). In some cases, the discovery of epistasis may merely be an artifact of using an incorrect null model (Kupper and Hogan 1978). It may be necessary to represent “independence” differently, resulting in different statistical measures of interactions, for different pairs of genes depending on their functions.Previous studies have suggested that different pairs of loci may have different modes of interaction and have attempted to subclassify genetic interactions into regulatory hierarchies and mutually exclusive “interaction subtypes” to elucidate underlying biological properties (Avery and Wasserman 1992; Drees et al. 2005; St. Onge et al. 2007). We suggest that epistatic relationships can be divided into several subtypes, or forms, corresponding to the aforementioned definitions of independence. As a particular gene pair may deviate from independence according to several criteria, we do not claim that these subtypes are necessarily mutually exclusive. We attempt to select the most likely epistatic subtype that is the best statistical representation of the relationship between two genes. To further subclassify interactions, epistasis among deleterious mutations can take one of two commonly used forms: positive (equivalently alleviating, antagonistic, or buffering) epistasis, where the phenotype of the double mutant is less severe than expected under independence, and negative (equivalently aggravating, synergistic, or synthetic), where the phenotype is more severe than expected (Segre et al. 2005; Collins et al. 2006; Desai et al. 2007; Mani et al. 2008).Another objective of such distinctions is to reduce the bias of the estimator of the epistatic parameter (ɛ), which measures the extent and direction of epistasis for a given gene pair. Mani et al. (2008), assuming that the overall distribution of ɛ should be centered around 0, find that inaccurately choosing a definition of independence can result in increased bias when estimating ɛ. For example, using the minimum definition results in the most severe bias when single mutants have moderate fitness effects, and the additive definition results in the largest positive bias when at least one gene has an extreme fitness defect (Mani et al. 2008). Therefore, it is important to select an optimal estimator for ɛ for each pair of genes from among the subtypes of epistatic interactions.Epistasis may be important to consider in genomic association studies, as a gene with a weak main effect may be identified only through its interaction with another gene or other genes (Frankel and Schork 1996; Culverhouse et al. 2002; Moore 2003; Cordell 2009; Moore and Williams 2009). Epistasis has also been studied extensively in the context of the evolution of sex and recombination. The mutational deterministic hypothesis proposes that the evolution of sex and recombination would be favored by negative epistatic interactions (Feldman et al. 1980; Kondrashov 1994); many other studies have also studied the importance of the form of epistasis (Elena and Lenski 1997; Otto and Feldman 1997; Burch and Chao 2004; Keightley and Otto 2006; Desai et al. 2007; MacCarthy and Bergman 2007). Indeed, according to Mani et al. (2008, p. 3466), “the choice of definition [of epistasis] alters conclusions relevant to the adaptive value of sex and recombination.”Given fitness data from single and double mutants in haploid organisms, we implement a likelihood method to determine the subtype that is the best statistical representation of the epistatic interaction for pairs of genes. We use maximum-likelihood estimation and the Bayesian information criteria (BIC) (Schwarz 1978) with a likelihood-ratio test to select the most appropriate null or epistatic model for each putative interaction. We conduct extensive simulations to gauge the performance of our method and demonstrate that it performs reasonably well under various interaction scenarios. We apply our method to two data sets with fitness measurements obtained from yeast (Jasnos and Korona 2007; St. Onge et al. 2007), whose authors assume only multiplicative epistasis for all interactions. By examining functional links and experimentally validated interactions among epistatic pairs, we demonstrate that our results are biologically meaningful. Studying a random selection of genes, we find that minimum epistasis is more prevalent than both additive and multiplicative epistasis and that the overall distribution of ɛ is not significantly different from zero (as Jasnos and Korona 2007 suggest). For genes in a particular pathway, we advise selecting among fewer epistatic subtypes. We believe that our method of epistatic subtype classification will aid in understanding genetic interactions and their properties.

St. Onge et al. (2007) data set:

St. Onge et al. (2007) examined 26 nonessential genes known to confer resistance to MMS, constructed double-deletion strains for 323 double-mutant strains (all but two of the total possible pairs), and assumed the multiplicative form of epistasis for all interactions (see Methods: Analysis of experimental data). Following these authors, we focus on single- and double-mutant fitnesses measured in the presence of MMS. (For results in the absence of MMS, see File S1 and File S1_2.)Using the resampling method described in Analysis of experimental data and File S1, 222 gene pairs pass the cutoff of having epistasis inferred in at least 900 of 1000 replicates. This does not include 5 synthetic lethal gene pairs. Hypothesis testing and a multiple-testing procedure (for 222 simultaneous hypotheses) are necessary to determine the final epistatic pairs.To select one among the three multiple-testing procedures, we follow St. Onge et al. (2007) and examine gene pairs that share specific functional links (see Analysis of experimental data). The Bonferroni method is likely too conservative, yielding only 25 significantly epistatic pairs with only one functional link among them; alternatively, the pFDR procedure appears to be too lenient in rejecting independence for all 222 pairs. Therefore, we use the FDR procedure (although the number of functional links is not significant) and detect 193 epistatic pairs, of which 5 (2.6%) are synthetic lethals, 19 (9.8%) have additive epistasis, 33 (17.1%) have multiplicative epistasis, and 136 (70.5%) have minimum epistasis (File S1_1). We find 29 gene pairs with positive (alleviating) epistasis and 159 gene pairs with negative (aggravating) epistasis.

TABLE 2

Summary of gene pairs with the indicated epistatic subtypes, inferred using the FDR procedure with the BIC method that considers all three epistatic subtypes and their corresponding null models
Epistatic subtypeStudy SStudy J
All193 (100%)352 (100%)
= −0.060 = −0.001
= −0.096 = −0.059
Additive19 (9.8%)35 (9.9%)
= 0.115* = 0.193***
= 0.131 = 0.188
Multiplicative33 (17.1%)63 (17.9%)
= 0.048 = 0.017
= −0.166 = −0.115
Minimum136 (70.5%)254 (72.2%)
= −0.111*** = −0.032**
= −0.091 = −0.065
Open in a separate windowNumbers are the counts of each type, and percentages are given of the total number of epistatic pairs. The mean () and median () of the epistatic parameter (ɛ) are given for each subtype, with “*” indicating that the mean of ɛ is significantly different from 0 (*, P-value ≤0.05; **, P-value ≤0.01; ***, P-value ≤0.001). Study S refers to the St. Onge et al. (2007) data set, and study J refers to the Jasnos and Korona (2007) data set. (For study S, five of the epistatic pairs are synthetic lethals and are not shown; as a result, percentages do not sum to 100%.)To further validate the use of our method and the FDR procedure, we assess by Fisher''s exact test the significance of an enrichment of both Biological Process and all GO Slim term links among epistatic pairs, neither of which are significant (Gene Ontology Consortium 2000; www.yeastgenome.org; Stark et al. 2006); Table S4]. Although some of the previously unidentified interactions that we identify could be false positives, many are likely to be new discoveries.

TABLE 3

Comparison of validation measures for each data set for different variations of the FDR and BIC procedures, considering only a subset of epistatic subtypes with their corresponding null models: all epistatic subtypes (A, P, and M); only the additive and multiplicative subtypes (A and P); and only the additive (A), only the multiplicative (P), or only the minimum (M) subtype (see text for details)
Subtypes considered in BIC procedure
A, P, MA, PAPM
Study J
No. found (636)352273263231329
Functional links (25)19 (0.0255)*13 (0.2320)11 (0.4689)10 (0.4227)15 (0.2619)
GO Slim terms (Biological Process) (115)69 (0.1573)50 (0.4874)55 (0.0736)44 (0.3534)68 (0.04902)*
GO Slim terms (all) (369)224 (0.0009)*172 (0.01654)*160 (0.1297)146 (0.0273)*213 (0.0003)*
Experimentally identified (3)32123
Study S
No. found (323)193192247171243
Functional links (36)21 (0.6450)29 (0.0041)*34 (0.0031)*29 (0.0003)*24 (0.9256)
GO Slim terms (Biological Process) (283)174 (0.0657)174 (0.03656)*223 (0.0010)*153 (0.1825)213 (0.5534)
GO Slim terms (all) (307)185 (0.2866)182 (0.6926)237 (0.1472)162 (0.6997)231 (0.5908)
Experimentally identified (29)1722242321
Open in a separate windowNumbers in parentheses indicate P-values by Fisher''s exact test. “*” indicates significance. Study J refers to the Jasnos and Korona (2007) data set, and study S refers to the St. Onge et al. (2007) data set measured in the presence of MMS. Numbers in parentheses indicate the total number of tested pairs and the total number of each type of link found in each complete data set.The epistatic subtypes we consider are not necessarily mutually exclusive. To more fully assess the assumptions of our method, we also consider several of the possible subsets of the epistatic subtypes (and their corresponding null models) in our procedure. As the minimum epistatic subtype was the most frequently selected in this data set, we first do not include the minimum null model or the minimum epistatic model in our procedure (i.e., we select from among four rather than six models for a pair; Table S4). However, there are a significant number of epistatic pairs with functional links only when the minimum epistatic subtype is not included (also see Table S4 and Table S5). It is not immediately clear which epistatic subtypes are the most appropriate for these data, although including the minimum subtype may not be appropriate (Mani et al. 2008) (see discussion).Although it may be best to consider fewer epistatic subtypes for this specific data set, we report our results including all three epistatic subtypes and their corresponding null models (St. Onge et al. (2007), although we identify 105 epistatic pairs not identified by the original authors (Figure S4, Table S4). St. Onge et al. (2007) find that epistatic pairs with a functional link have a positively shifted distribution of epistasis. We find no such shift in epistasis values (Figure S5). We also demonstrate [described in application to simulated data: Bias and variance of the epistatic parameter (ɛ)] that our method seems to reduce bias of the epistatic parameter (ɛ) (Table S3).] When considering only a subset of the epistatic subtypes, however, we find to be positive and significantly different from zero (results not shown). See File S1, Figure S6, and Figure S7 for additional discussion of the epistatic pairs we identify.

Jasnos and Korona (2007) data set:

The Jasnos and Korona (2007) data set included 758 yeast gene deletions known to cause growth defects and reports fitnesses of only a sparse subset of all possible gene pairs [≈0.2% of the possible pairwise genotypes, or 639 pairs of ]. Because the authors do not identify epistatic pairs in a hypothesis-testing framework, we cannot explicitly compare our conclusions with theirs.To validate our method, we examine gene pairs that have specific functional links (see methods: Analysis of experimental data). When defining a functional link using GO terms (Gene Ontology Consortium 2000) with <30 genes associated with them, only 1 of 639 tested gene pairs has a functional link. Raising the threshold of associated genes to 50 and 100, the number of tested pairs with functional links rises only to 3 and 9, respectively. Because of the large number of random genes and the sparse number of gene pairs in this data set, we follow Tong et al. (2004) and select GO terms that have associated with them ≤200 genes. Twenty-five of 639 tested pairs then have a functional link.Only the FDR multiple-testing procedure results in a significant enrichment of functional links among epistatic pairs (File S1). With the FDR procedure we find 352 significant epistatic pairs, of which 35 (9.9%) have additive epistasis, 63 (17.9%) have multiplicative epistasis, and 254 (72.2%) have minimum epistasis (File S1_3). These proportions of inferred subtypes suggest that the authors'' original restriction to multiplicative epistasis may be inappropriate. We find 141 gene pairs with positive epistasis and 211 gene pairs with negative epistasis.We do not find a significant number of epistatic pairs with shared GO Slim Biological Process terms (see Analysis of experimental data), but do when considering all shared GO Slim terms (St. Onge et al. (2007) data set, we also consider some of the possible subsets of the three epistatic subtypes (and their corresponding null models) in our model selection procedure (Table S5). In contrast to the St. Onge et al. (2007) data set, using all three epistatic subtypes results in a significant number of epistatic pairs with functional links; this measure is not significant when using any of the other subsets of the subtypes. This suggests that our proposed method with three epistatic subtypes may indeed be the most appropriate for data sets with randomly selected genes.We examined the distribution of the estimated values of the epistatic parameter (ɛ) for all pairs with significant epistasis. Jasnos and Korona (2007), in assuming only multiplicative epistasis, conclude that epistasis is predominantly positive. However, we find that the estimated mean of epistasis is not significantly different from zero (two-sided t-test, P-value = 0.9578; Figure 1 and File S1.Open in a separate windowFigure 1.—Distribution of the epistasis values (ɛ) for significant epistatic pairs in the Jasnos and Korona (2007) data set, determined using the FDR procedure and the BIC method including all three epistatic subtypes and their corresponding null models. Mean of ɛ is −0.0009, with a standard deviation of 0.3177; median value is −0.0587. A similar plot is shown in Figure 3 of Jasnos and Korona (2007).  相似文献   

15.
A Problem With the Correlation Coefficient as a Measure of Gene Expression Divergence     
Vini Pereira  David Waxman  Adam Eyre-Walker 《Genetics》2009,183(4):1597-1600
The correlation coefficient is commonly used as a measure of the divergence of gene expression profiles between different species. Here we point out a potential problem with this statistic: if measurement error is large relative to the differences in expression, the correlation coefficient will tend to show high divergence for genes that have relatively uniform levels of expression across tissues or time points. We show that genes with a conserved uniform pattern of expression have significantly higher levels of expression divergence, when measured using the correlation coefficient, than other genes, in a data set from mouse, rat, and human. We also show that the Euclidean distance yields low estimates of expression divergence for genes with a conserved uniform pattern of expression.IT is now possible to measure the expression levels of thousands of genes in multiple tissues at multiple times. This has led to investigations into the evolution of gene expression and how the pattern of expression changes on a genomic scale. In some analyses, the evolution of expression is considered only within one tissue, but in many studies the evolution across multiple tissues is investigated. In this latter case, the evolution of an expression profile—a vector of expression levels of a gene across several tissues—is considered.Several different statistics have been proposed to measure the divergence between gene expression profiles. The two most popular measures are the Euclidean distance (Jordan et al. 2005; Kim et al. 2006; Yanai et al. 2006; Urrutia et al. 2008) and Pearson''s correlation coefficient (Makova and Li 2003; Huminiecki and Wolfe 2004; Yang et al. 2005; Kim et al. 2006; Liao and Zhang 2006a,b; Xing et al. 2007; Urrutia et al. 2008). The correlation coefficient is often subtracted from one, so that the statistic varies from zero, when there has been no expression divergence, to a maximum of two; we refer to this statistic as the Pearson distance. Here we describe a significant shortcoming of the Pearson distance that is not shared by the Euclidean distance.To investigate properties of these two measures of expression divergence, we compiled a data set of 2859 orthologous genes from human, mouse, and rat for which we had microarray expression data from nine homologous tissues: bone marrow, heart, kidney, large intestine, pituitary, skeletal muscle, small intestine, spleen, and thymus). The expression data for rat came from Walker et al. (2004), the mouse data from Su et al. (2004), and the human data from Ge et al. (2005). Each tissue experiment had two replicates in mouse, a varying number of replicates in rat, and one in humans; some genes were also matched by multiple probe sets. To obtain an average across experiments and probe sets we processed the data as follows:
  1. Raw CEL files of gene expression levels were obtained from the NCBI Gene Expression Omnibus database (http://www.ncbi.nlm.nih.gov/projects/geo/).
  2. The results from the mouse, rat, and human arrays were normalized separately using both the MAS5 (Affymetrix 2001) and the RMA algorithms (Irizarry et al. 2003) as implemented in Bioconductor (Gentleman et al. 2004). The results are qualitatively similar for the two normalization procedures, although recent analyses suggest that MAS5 normalization is generally better (Ploner et al. 2005; Lim et al. 2007).
  3. The expression of each gene within a tissue was averaged across experiments and probe sets.
We computed expression distances (ED) between orthologous gene expression profiles, for each of the three species comparisons, rat–mouse, rat–human, and mouse–human, according to the two different distance metrics, the Euclidean distance and the Pearson distance:(1)Here xij is the expression level of the gene under consideration in species i in tissue j, and is the average expression level of the gene in species i across tissues. Expression levels are known in a total of k tissues.Because expression levels are measured on different microarray platforms in the three species, we compute relative abundance (RA) values, before calculating the Euclidean distance (Liao and Zhang 2006a). The RA is the expression of a gene in a particular tissue divided by the sum of the expression values of that gene across all tissues. We calculated RA values to remove “probe” effects (the tendency for a gene to bind its probe set on one platform more efficiently than on another platform). Because of probe effects it is not easy to distinguish absolute changes in expression and differences in binding efficiency. Calculating RA values removes this problem from the Euclidean distance. Pearson''s distance does not change under such a rescaling and so this is unnecessary.In some analyses the logarithm of the expression or RA values are used (e.g., Makova and Li 2003; Kim et al. 2006; Xing et al. 2007), and in others the expression values are used without this transformation (e.g., Huminiecki and Wolfe 2004; Jordan et al. 2005; Yang et al. 2005; Liao and Zhang 2006a,b; Yanai et al. 2006; Urrutia et al. 2008). We calculated both the Pearson and the Euclidean distances on log-transformed and untransformed expression values. The results are qualitatively similar so here we present only the results obtained using the logarithm of the expression or RA values.It is natural to expect the two measures of expression divergence to be positively correlated with one another; however, the Euclidean and Pearson distances are almost completely uncorrelated (MAS5 normalization, mouse–rat correlation coefficient = 0.06, human–rat r = 0.13, human–mouse r = 0.10; RMA normalization, mouse–rat correlation coefficient = −0.12, human–rat r = −0.00, human–mouse r = −0.08; Figure 1). This could, plausibly, be because the two statistics measure different aspects of divergence. However, irrespective of this, there is a potential problem associated with the Pearson distance. Imagine that we have a gene that is expressed at identical levels in all tissues in two species (i.e., expression levels are uniform between tissues and also between species). We quite reasonably assume that measured expression levels contain noise. Thus each measured expression level (xij) is the sum of the (assumed) uniform expression level and an independent random number representing noise. In this case there is no real divergence in the expression profile between the species. However, the two measures of divergence may differ greatly in this case. The Euclidean distance reflects only the noise present in the data and hence will be small if the noise is small. By contrast, the Pearson distance will have a value close to 1 since the second term in PeaD in Equation 1 will be close to zero, reflecting the fact that the noise components of different expression levels are independent. Thus the Pearson distance will give the impression that expression divergence is great, but all this apparent divergence is noise. This will be a problem with Pearson''s distance whenever measurement error is of the same magnitude as the differences in expression between tissues. This will therefore tend to be a problem for lowly expressed genes, where measurement error can be large relative to the true value.Open in a separate windowFigure 1.—The correlation between the Euclidean and Pearson distances for (a) mouse–rat, (b) human–rat, and (c) human–mouse. Only the results from MAS5 normalization are shown; qualitatively similar results were obtained with RMA.The above example is unrealistic because real gene expression profiles are rarely perfectly uniform. To investigate whether this shortcoming of the Pearson distance is a problem in real data sets, we determined genes with a relatively uniform pattern of expression in all three species considered above. To do this we computed the entropy of a gene''s expression, which is a measure of uniformity in expression across tissues (Schug et al. 2005): the higher the value of the entropy, the more uniform is the expression. We calculated the entropy for each gene in each of the three species, averaged these across species, and then took those genes in the upper quartile of mean entropy values as a data set of genes with a relatively conserved pattern of uniform expression.It is natural to expect those genes with a conserved uniform pattern of expression to have relatively low expression divergence; however, on average these genes have significantly higher Pearson distances than other genes (Figure 2; supporting information, Figure S1 and Figure S2). By contrast, the Euclidean distance shows the pattern one would anticipate; all of the conserved uniform genes have low expression divergence. It therefore seems likely that the Pearson distance is sensitive to measurement error and hence may not be a good measure of expression divergence.Open in a separate windowFigure 2.—The distribution of expression divergence values for those genes with a uniform pattern of expression that is conserved across species vs. the distribution for all genes for (a) Pearson and (b) Euclidean distances for mouse–rat. We present similar values for human–mouse and human–rat in Figure S1 and Figure S2. Only the results from MAS5 normalization are shown; qualitatively similar results were obtained with RMA.

TABLE 1

The median expression divergence for genes that have a conserved uniform pattern of expression (upper quartile of mean entropy values) vs. all other genes
Data setStatisticConserved uniform genesOther genesWilcoxon test P-value
MAS5 normalization
    Mouse–ratEuclidean1.662.79<10−15
Pearson0.700.47<10−15
    Human–mouseEuclidean1.673.13<10−15
Pearson0.780.58<10−15
    Human–ratEuclidean1.833.21<10−15
Pearson0.780.58<10−15
RMA normalization
    Mouse–ratEuclidean0.591.40<10−15
Pearson0.820.38<10−15
    Human–mouseEuclidean0.591.58<10−15
Pearson0.810.48<10−15
    Human–ratEuclidean0.581.55<10−15

Pearson
0.73
0.50
<10−15
Open in a separate windowWe note that there are two additional advantages of the Euclidean distance. First, it can take into account differences in the absolute level of expression if those data are available, either because the method of assay allows this, for example, if ESTs, SAGE, sequencing, or RNA-Seq data are used, or because expression in the two species has been assessed on the same platform using probes that are conserved between the two species. Second, the square of the Euclidean distance is expected to increase linearly with time. Khaitovich et al. (2004) have previously shown that the squared difference in log expression level increases linearly with time under a Brownian motion model of gene expression evolution. It is therefore expected that the squared Euclidean distance will increase with time since the squared Euclidean distance is the sum of the squared differences across tissues. We prove this in File S1; we also show that this linearity holds, approximately, when relative abundance values are used (see also Pereira et al. 2009).  相似文献   

16.
Reciprocal Silencing,Transcriptional Bias and Functional Divergence of Homeologs in Polyploid Cotton (Gossypium)     
Bhupendra Chaudhary  Lex Flagel  Robert M. Stupar  Joshua A. Udall  Neetu Verma  Nathan M. Springer  Jonathan F. Wendel 《Genetics》2009,182(2):503-517
  相似文献   

17.
Tissue-Specific Expression Patterns of Arabidopsis NF-Y Transcription Factors Suggest Potential for Extensive Combinatorial Complexity          下载免费PDF全文
Nicholas Siefers  Kristen K. Dang  Roderick W. Kumimoto  William Edwards Bynum  IV  Gregory Tayrose  Ben F. Holt  III 《Plant physiology》2009,149(2):625-641
  相似文献   

18.
Stable Transcription Activities Dependent on an Orientation of Tam3 Transposon Insertions into Antirrhinum and Yeast Promoters Occur Only within Chromatin     
Takako Uchiyama  Kaien Fujino  Takashi Ogawa  Akihito Wakatsuki  Yuji Kishima  Tetsuo Mikami  Yoshio Sano 《Plant physiology》2009,151(3):1557-1569
  相似文献   

19.
Exciting Prospects for Precise Engineering of Caenorhabditis elegans Genomes with CRISPR/Cas9     
Christian Fr?kj?r-Jensen 《Genetics》2013,195(3):635-642
  相似文献   

20.
Arabidopsis LON2 Is Necessary for Peroxisomal Function and Sustained Matrix Protein Import     
Matthew J. Lingard  Bonnie Bartel 《Plant physiology》2009,151(3):1354-1365
Relatively little is known about the small subset of peroxisomal proteins with predicted protease activity. Here, we report that the peroxisomal LON2 (At5g47040) protease facilitates matrix protein import into Arabidopsis (Arabidopsis thaliana) peroxisomes. We identified T-DNA insertion alleles disrupted in five of the nine confirmed or predicted peroxisomal proteases and found only two—lon2 and deg15, a mutant defective in the previously described PTS2-processing protease (DEG15/At1g28320)—with phenotypes suggestive of peroxisome metabolism defects. Both lon2 and deg15 mutants were mildly resistant to the inhibitory effects of indole-3-butyric acid (IBA) on root elongation, but only lon2 mutants were resistant to the stimulatory effects of IBA on lateral root production or displayed Suc dependence during seedling growth. lon2 mutants displayed defects in removing the type 2 peroxisome targeting signal (PTS2) from peroxisomal malate dehydrogenase and reduced accumulation of 3-ketoacyl-CoA thiolase, another PTS2-containing protein; both defects were not apparent upon germination but appeared in 5- to 8-d-old seedlings. In lon2 cotyledon cells, matrix proteins were localized to peroxisomes in 4-d-old seedlings but mislocalized to the cytosol in 8-d-old seedlings. Moreover, a PTS2-GFP reporter sorted to peroxisomes in lon2 root tip cells but was largely cytosolic in more mature root cells. Our results indicate that LON2 is needed for sustained matrix protein import into peroxisomes. The delayed onset of matrix protein sorting defects may account for the relatively weak Suc dependence following germination, moderate IBA-resistant primary root elongation, and severe defects in IBA-induced lateral root formation observed in lon2 mutants.Peroxisomes are single-membrane-bound organelles found in most eukaryotes. Peroxin (PEX) proteins are necessary for various aspects of peroxisome biogenesis, including matrix protein import (for review, see Distel et al., 1996; Schrader and Fahimi, 2008). Most matrix proteins are imported into peroxisomes from the cytosol using one of two targeting signals, a C-terminal type 1 peroxisome-targeting signal (PTS1) or a cleavable N-terminal type 2 peroxisome-targeting signal (PTS2) (Reumann, 2004). PTS1- and PTS2-containing proteins are bound in the cytosol by soluble matrix protein receptors, escorted to the peroxisome membrane docking complex, and translocated into the peroxisome matrix (for review, see Platta and Erdmann, 2007). Once in the peroxisome, many matrix proteins participate in metabolic pathways, such as β-oxidation, hydrogen peroxide decomposition, and photorespiration (for review, see Gabaldon et al., 2006; Poirier et al., 2006).In addition to metabolic enzymes, several proteases are found in the peroxisome matrix. Only one protease, DEG15/Tysnd1, has a well-defined role in peroxisome biology. The rat Tysnd1 protease removes the targeting signal after PTS2-containing proteins enter the peroxisome and also processes certain PTS1-containing β-oxidation enzymes (Kurochkin et al., 2007). Similarly, the Arabidopsis (Arabidopsis thaliana) Tysnd1 homolog DEG15 (At1g28320) is a peroxisomal Ser protease that removes PTS2 targeting signals (Helm et al., 2007; Schuhmann et al., 2008).In contrast with DEG15, little is known about the other eight Arabidopsis proteins that are annotated as proteases in the AraPerox database of putative peroxisomal proteins (Reumann et al., 2004; Carter et al., 2004; Shimaoka et al., 2004), which, in combination with the minor PTS found in both of these predicted proteases (Reumann, 2004), suggests that these enzymes may not be peroxisomal. Along with DEG15, only two of the predicted peroxisomal proteases, an M16 metalloprotease (At2g41790), which we have named PXM16 for peroxisomal M16 protease, and a Lon-related protease (At5g47040/LON2; Ostersetzer et al., 2007), are found in the proteome of peroxisomes purified from Arabidopsis suspension cells (Eubel et al., 2008). DEG15 and LON2 also have been validated as peroxisomally targeted using GFP fusions (Ostersetzer et al., 2007; Schuhmann et al., 2008).

Table I.

Putative Arabidopsis proteases predicted or demonstrated to be peroxisomal
AGI IdentifierAliasProtein ClassT-DNA Insertion AllelesPTSLocalization EvidenceLocalization References
At1g28320DEG15PTS2-processing proteaseSALK_007184 (deg15-1)SKL>aGFPReumann et al., 2004; Helm et al., 2007; Eubel et al., 2008; Schuhmann et al., 2008)
Proteomics
Bioinformatics
At2g41790PXM16Peptidase M16 family proteinSALK_019128 (pxm16-1)PKL>bProteomicsReumann et al., 2004, 2009; Eubel et al., 2008)
SALK_023917 (pxm16-2)Bioinformatics
At5g47040LON2Lon protease homologSALK_128438 (lon2-1)SKL>aGFPReumann et al., 2004, 2009; Ostersetzer et al., 2007; Eubel et al., 2008)
SALK_043857 (lon2-2)Proteomics
Bioinformatics
At2g18080Ser-type peptidaseSALK_020628SSI>cBioinformatics(Reumann et al., 2004)
SALK_102239
At2g35615Aspartyl proteaseSALK_090795ANL>bBioinformatics(Reumann et al., 2004)
SALK_036333
At3g57810Ovarian tumor-like Cys proteaseSKL>aBioinformatics(Reumann et al., 2004)
At4g14570Acylaminoacyl-peptidase proteinCKL>bBioinformatics (peroxisome)(Reumann et al., 2004; Shimaoka et al., 2004)
Proteomics (vacuole)
At4g20310Peptidase M50 family proteinRMx5HLdBioinformatics(Reumann et al., 2004)
At4g36195Ser carboxypeptidase S28 familySSM>bBioinformatics (peroxisome)(Carter et al., 2004; Reumann et al., 2004)





Proteomics (vacuole)

Open in a separate windowaMajor PTS1 (Reumann, 2004).bMinor PTS1 (Reumann, 2004).cValidated PTS1 (Reumann et al., 2007).dMinor PTS2 (Reumann, 2004).PXM16 is the only one of the nine Arabidopsis M16 (pitrilysin family) metalloproteases (García-Lorenzo et al., 2006; Rawlings et al., 2008) containing a predicted PTS. M16 subfamilies B and C contain the plastid and mitochondrial processing peptidases (for review, see Schaller, 2004), whereas PXM16 belongs to M16 subfamily A, which includes insulin-degrading peptidases (Schaller, 2004). A tomato (Solanum lycopersicum) M16 subfamily A protease similar to insulin-degrading enzymes with a putative PTS1 was identified in a screen for proteases that cleave the wound response peptide hormone systemin (Strassner et al., 2002), but the role of Arabidopsis PXM16 is unknown.Arabidopsis LON2 is a typical Lon protease with three conserved domains: an N-terminal domain, a central ATPase domain in the AAA family, and a C-terminal protease domain with a Ser-Lys catalytic dyad (Fig. 1A; Lee and Suzuki, 2008). Lon proteases are found in prokaryotes and in some eukaryotic organelles (Fig. 1C) and participate in protein quality control by cleaving unfolded proteins and can regulate metabolism by controlling levels of enzymes from many pathways, including cell cycle, metabolism, and stress responses (for review, see Tsilibaris et al., 2006). Four Lon homologs are encoded in the Arabidopsis genome; isoforms have been identified in mitochondria, plastids, and peroxisomes (Ostersetzer et al., 2007; Eubel et al., 2008; Rawlings et al., 2008). Mitochondrial Lon protesases are found in a variety of eukaryotes (Fig. 1A) and function both as ATP-dependent proteases and as chaperones promoting protein complex assemblies (Lee and Suzuki, 2008). LON2 is the only Arabidopsis Lon isoform with a canonical C-terminal PTS1 (SKL-COOH; Ostersetzer et al., 2007) or found in the peroxisome proteome (Eubel et al., 2008; Reumann et al., 2009). Functional studies have been conducted with peroxisomal Lon isoforms found in the proteome of peroxisomes purified from rat hepatic cells (pLon; Kikuchi et al., 2004) and the methylotrophic yeast Hansenula polymorpha (Pln; Aksam et al., 2007). Rat pLon interacts with β-oxidation enzymes, and a cell line expressing a dominant negative pLon variant has decreased β-oxidation activity, displays defects in the activation processing of PTS1-containing acyl-CoA oxidase, and missorts catalase to the cytosol (Omi et al., 2008). H. polymorpha Pln is necessary for degradation of a misfolded, peroxisome-targeted version of dihydrofolate reductase and for degradation of in vitro-synthesized alcohol oxidase in peroxisomal matrix extracts, but does not contribute to degradation of peroxisomally targeted GFP (Aksam et al., 2007).Open in a separate windowFigure 1.Diagram of LON2 protein domains, gene models for LON2, PXM16, DEG15, PED1, PEX5, and PEX6, and phylogenetic relationships of LON family members. A, Organization of the 888-amino acid LON2 protein. Locations of the N-terminal domain conserved among Lon proteins, predicted ATP-binding Walker A and B domains (black circles), active site Ser (S) and Lys (K) residues (asterisks), and the C-terminal Ser-Lys-Leu (SKL) peroxisomal targeting signal (PTS1) are shown (Lee and Suzuki, 2008). B, Gene models for LON2, PXM16, DEG15, PED1, PEX5, and PEX6 and locations of T-DNA insertions (triangles) or missense alleles (arrows) used in this study. Exons are depicted by black boxes, introns by black lines, and untranslated regions by gray lines. C, Phylogenetic relationships among LON homologs. Sequences were aligned using MegAlign (DNAStar) and the ClustalW method. The PAUP 4.0b10 program (Swofford, 2001) was used to generate an unrooted phylogram from a trimmed alignment corresponding to Arabidopsis LON2 residues 400 to 888 (from the beginning of the ATPase domain to the end of the protein). The bootstrap method was performed for 500 replicates with distance as the optimality criterion. Bootstrap values are indicated at the nodes. Predicted peroxisomal proteins have C-terminal PTS1 signals in parentheses and are in light-gray ovals. Proteins in the darker gray oval have N-terminal extensions and include mitochondrial and chloroplastic proteins. Sequence identifiers are listed in Supplemental Table S2.In this work, we examined the roles of several putative peroxisomal proteases in Arabidopsis. We found that lon2 mutants displayed peroxisome-deficient phenotypes, including resistance to the protoauxin indole-3-butyric acid (IBA) and age-dependent defects in peroxisomal import of PTS1- and PTS2-targeted matrix proteins. Our results indicate that LON2 contributes to matrix protein import into Arabidopsis peroxisomes.  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