首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The transition density for multiple neutral alleles   总被引:2,自引:0,他引:2  
For a single genetic locus with multiple alleles, Littler and Fackerall (1975, Biometrics 31, 117–123) found the transition density of the allelic proportions in the case of no mutation and no selection. Their method is found to be suitable for a model in which symmetric mutation is allowed between alleles. The sampling probabilities and moments are found. Watterson's (1977, Genetics 85, 789–814) test for selection is found to be non-robust for non-stationary populations.  相似文献   

2.
The consequences of preferential mating in the presence of partial assortative and sexual selection mechanisms are ascertained for a two-allele one-locus trait involving two phenotype classes C1 = {all homozygotes} and C2 = {heterozygotes}. Relevant biological cases may include Burley (1977, Proc. Nat. Acad. Sci. USA 74, 3476–3479), Wilbur et al. (1978, Evolution 32, 264–270), and Singh and Zouros (1978, Evolution 32, 342–353). When the preference rate for the heterozygote exceeds that for homozygotes, it is established that the unique stable state is the central Hardy-Weinberg equilibrium. The rate of approach is faster with sexual selection than for the corresponding model of assortative mating. When the preference rates favor the homozygotes then in this symmetric model of sexual selection two asymmetric Hardy-Weinberg polymorphisms can evolve, and which succeeds depends on initial conditions. The models are also analyzed with natural selection acting on phenotypes superimposed on assortative mating. In this case we can have up to three coexisting stable states involving both fixation alternatives and a central polymorphism. The corresponding model with sexual selection maintains either the central equilibrium as in assortative mating or two asymmetric polymorphic equilibria.  相似文献   

3.
Competition in a coarse-grained heterogeneous environment is considered. It is assumed that each individual adjust has position within the macrohabitat so as to try to maximize a utility whose value is a function of position within the habitat and of the population density at that position. Robust qualitative conclusions concerning the resulting population behavior relate niche breadth, niche shift, and resource partitioning to changes in population numbers. In particular populations avoid competition by reducing coarse-grained niche overlap so that overlap is a measure not of the degree of competition but of its absence. Specific models of exploitation or interference competition give specific quantitative results which are summarized. Competition equations are derived from underlying models which imply Lotka-Volterra equations in a fine-grained environment (MacArthur, R., and Levins, R. 1967. Amer. Natur.101, 377–385) but which are nonlinear in per capita growth in a coarse-grained one. The findings are relevant to theoretical studies of competition, limiting similarity, species packing, and the evolution of niche position as well as to practical problems of data analysis and resource management. The extension to predator-prey interactions is outlined.  相似文献   

4.
Plants of the genus Pistacia (Anacardiaceae) serve as obligate hosts for a group of specialized gall-forming aphids (Homoptera: Fordinae). The aphids regularly migrate between the Pistacia (primary) host plants and the roots of non-specific grasses and cereals (secondary hosts). Gall density varies considerably between trees and sites. The intimate relationships between the aphids and their primary host, the natural variation of host susceptibility, and the heterogeneous geographical environment may promote local adaptation and deme formation in the aphid populations. Indeed, previous analyses of the genetic structure of the aphid Baizongia pistaciae, which forms large galls on the deciduous P. palaestina trees, suggested deme formation (Martinez et al. 2005). In this study, we analyzed the genetic structure of the B. pistaciae population at eight sites and 78 trees throughout Israel and a single population in Turkey, using two molecular markers (AFLP fingerprints and COI sequencing). The genetic distance between the Israeli populations was found to be low (D = 0.01–0.02), and there was no genetic differentiation found between any population pairs. In five of the Israeli populations, we also compared the genetic identity between aphids forming galls on the same tree and between galls on neighboring trees. The analysis indicated that the genetic identity of different galls within a tree resembles the correspondence between trees within a population. Our results showed no indication of deme formation or any hierarchical genetic substructuring within B. pistaciae populations in Israel. The extensive gene flow between aphid colonies may be explained by their dispersal abilities and the potential bridging role of the secondary hosts.  相似文献   

5.
The preparation and crystal structures for three Cu(II) polynuclear, chain complexes with 2,3-bis(2-pyridyl)pyrazine (dpp) as bridging ligand are reported, [Cu(dpp)(H2O)2]n(NO3)2n·2n/3H2O (1), [Cu(dpp)(H2O)2]n(CF3SO3)2n (2), and [Cu(dpp)(H2O)2]n(BF4)2n·2nH2O (3). For the latter compound the crystal structure at four different temperatures have been studied. Variable-temperature magnetic susceptibility data and ESR measurements of 13 and of the related copper(II) chain [Cu(dpp)(H2O)2]n(ClO4)2n·2nH2O (4) (whose structure was previously reported) have been performed. Compounds 1 and 2 crystallize in the same trigonal space group, R c; 3 is orthorhombic, space group Pbca. Complexes 1 and 2 are built of linear dpp-bridged chains which extend along threefold screw axes. The copper atom has in each case an elongated octahedral geometry with pyrazine nitrogen atoms in axial positions. The prominent feature of the crystal packing is the supramolecular arrangement of six chains around a threefold inversion axis, creating channels housing the counter ions, and in the case of 1, also crystal water. In 3 the chain is zig–zag shaped and extends along a twofold screw axis. Counter ions and crystal water are situated in channels formed between four symmetry related chains. At room temperature (r.t.) the X-ray results show a copper ion with a compressed octahedral coordination geometry, pyrazine and pyridyl nitrogen atoms binding in equatorial and axial positions, respectively. The low temperature X-ray studies of 3 show significant changes in the copper coordination geometry, strongly suggesting that the apparent compressed geometry at r.t. is due to a dynamic Jahn–Teller distortion. The CuCu separations across the bridging dpp at r.t. are, 7.133(1) (1), 7.228(1) (2), 7.005(1) (3) and 7.002(2) Å (4). X-band ESR spectra of 1 and 2 exhibit the pattern of Cu(II) in elongated geometry (g>g>2.0), whereas those of 3 and 4 exhibit inverse (g>g>2.0) patterns with a shoulder in the perpendicular signal. Complexes 14 exhibit a Curie law behaviour with very weak intrachain antiferromagnetic coupling, the relevant magnetic parameters being J=−0.27 cm−1, g=2.11 for 1, J=−0.17 cm−1, g=2.09 for 2, J=−1.38 cm−1, g=2.15 for 3, and J=−1.36 cm−1, g=2.14 for 4 (the Hamiltonian being =−JSASB).  相似文献   

6.
Le Bras (Theor. Popul. Biol.2, 100–121, 1971) and Rogers (Demography11, 473–481, 1974), in two neglected papers, have generalized to the multisite case the Euler–Lotka renewal equation and demographic characteristics such as age structure and reproductive value. The purpose of this paper is twofold: first, to restate the multisite renewal equation in the matrix context; second, to derive results on age structure, net reproduction rate, generation time, and sensitivities, as generalizations of the one site case. The potential of this approach for population biology is illustrated using a model of a black-headed gullLarus ridibunduspopulation.  相似文献   

7.
An analytic model is developed to explore the relationship between gene flow, selection, and genetic drift. We assume that a single copy of a mutant allele appears in a finite, partially isolated population and allow for the effects of immigration, genic selection, and mutation on the frequency of the mutant. Our concern is with the distribution of the mutant's frequency before it either is lost from the population or emigrates. Before either of these events, the allele will be a “private allele” and would be found in only one of several populations in a larger collection. Slatkin [(1985) Evolution 39, 53–65] found several simple properties of private alleles in his simulations. We use the method developed by Karlin and Tavaré [(1980) Genet. Res. 37, 33–46; (1981a), Theor. Pop. Biol. 19, 187–214; (1981b) Theor. Pop. Biol. 19, 215–229] for a model similar to ours to obtain a diffusion equation with a “killing term” and obtain the mean and variance of the mutant's frequency and its expected frequency in samples of a specified size. There is only fair agreement between the analytic results from this model and those from Slatkin's (loc. cit.) simulations. The rescaling method used to obtain the results indicates that if emigration is relatively frequent, the distribution of rare alleles is governed largely by the balance between genetic drift and emigration, with selection, mutation, and immigration playing a lesser role.  相似文献   

8.
A population of constant size is subjected to mutation, such that each mutant is of a new allelic type. For the particular population model studied in this paper, the age of an allele, whose present frequency is known, is a random variable with distribution independent of the frequencies of other alleles. As a consequence of reversibility of the population process, the age of an allele, from the past to the present, has the same distribution as its time to extinction, from the present into the future. This verifies, and re-interprets, certain diffusion approximations found by Kimura and Ohta [Genetics 75, 199–212 (1973)] and Maruyama [Genet. Res. Cambridge 23, 137–143 (1974)].  相似文献   

9.
J. B. S. Haldane (Amer. Nat. 71, 337–349, 1937) argued that, in equilibrium populations, the effect of deleterious mutation on average fitness depends primarily on the mutation rate and is independent of the severity of the mutations. Specifically, the equilibrium population fitness is e−μH, where μH is the haploid genomic mutation rate. Here we extend Haldane's result to a variety of reproductive systems. Using an analysis based on the frequency of classes of individuals with a specified number of mutations, we show that Haldane's principle holds exactly for haploid sex, haploid apomixis, and facultative haploid sex. In the cases of diploid automixis with terminal fusion, diploid automixis with central fusion, and diploid selfing, Haldane's principle holds exactly for recessive mutations and approximately for mutations with some heterozygous effect. In the cases of K-ploid apomixis, diploid endomitosis, and haplodiploidy, we show that Haldane's principle holds exactly for recessive lethal mutations. In addition we extend Haldane's result to various mixtures of the above-mentioned reproductive systems. In the case of diploid out-crossing sexuals, we do not obtain an exact analytic result, but present arguments and computer simulations which show that Haldane's result extends to this case as well in the limit as the number of loci becomes large. Although diverse reproductive systems are equally fit at equilibrium, different reproductive systems harbor vastly different numbers of recessive genes at equilibrium and we provide estimates of these numbers. These different numbers of mutations may create transient selective pressures on individuals with reproductive systems different from that of the equilibrium population.  相似文献   

10.
Home range (HR) size and overlap, movements, and nest site use of fat dormice Glis glis (Linnaeus, 1766) were studied by live-trapping and radiotracking in the Bialowieza Forest in north-eastern Poland. The study was conducted during a mast year of oak Quercus robur (2001) and hornbeam Carpinus betulus (2002). Average HR size measured by 100% minimum convex polygons varied from 3.6 to 7.0 ha in males and from 0.55 to 0.76 ha in females. HR size in males was significantly larger in the hornbeam mast year than in the oak mast year, which was probably due to the lower energy value of hornbeam mast. Males had significantly larger HRs and used more nest sites than females. In contrast to females, HRs of males overlapped extensively. In the mating season, core areas of males overlapped significantly more and they shared their core areas with more males than in the post-mating season. During the mating season, males were found to share nest sites with other males or with a single female. Additionally, groups of 2–3 males were observed aggregating around a female during the mating season. Our data suggest that fat dormice have a promiscuous mating system where the females are territorial, and the non-territorial males directly compete for access to receptive females during the mating season.  相似文献   

11.
Restriction enzyme map for streptomycete plasmid pUC3   总被引:1,自引:0,他引:1  
A restriction enzyme map for the streptomycete plasmid pUC3 was constructed for the enzymes XhoI, EcoRI, HindIII, PstI, BamH-I, and BglII. The plasmid was isolated from Streptomyces sp. 3022a which produces chloramphenicol and has been referred to as S. venezuelae (Bewick et al., 1976 and Bewick and Williams, 1977, Microbios, 19, 27–35).  相似文献   

12.

Objective

To assess whether HIV surveillance data from pregnant women attending antenatal care (ANC) clinics in Zimbabwe represent infection levels in the general population.

Methods

HIV prevalence estimates from ANC surveillance sites in 2006 were compared with estimates from the corresponding Zimbabwe Demographic and Health Survey 2005–06 (ZDHS) clusters using geographic information systems.

Results

The ANC HIV prevalence estimate (17.9%, 95% CI 17.0%–18.8%) was similar to the ZDHS estimates for all men and women aged 15–49 years (18.1%, 16.9%–18.8%), for pregnant women (17.5%, 13.9%–21.9%), and for ANC attendees living within 30 km of ANC surveillance sites (19.9%, 17.1%–22.8%). However, the ANC surveillance estimate (17.9%) was lower than the ZDHS estimates for all women (21.1%, 19.7%–22.6%) and for women living within 30 km catchment areas of ANC surveillance sites (20.9%, 19.4%–22.3%). HIV prevalence in ANC sites classified as urban and rural was significantly lower than in sites classified as “other”.

Conclusions

Periodic population surveys can be used to validate ANC surveillance estimates. In Zimbabwe, ANC surveillance provides reliable estimates of HIV prevalence among men and women aged 15–49 years in the general population. Three classifications of ANC sites (rural/urban/other) should be used when generating national HIV estimates.  相似文献   

13.
A dynamical equation for the spatial distribution of competing species that contains a growth term and a dispersal term is analyzed under the condition that the dispersal rate is sufficiently rapid compared to the growth rate. With this assumption, the equation becomes analogous to the niche-partitioning theory of MacArthur and Levins (1967, Amer. Natur. 101, 377–385) and thus provides a link between the theory of local niche partitioning and the theory of regional habitat segregation. The formulation is applied to a two species system consisting of a specialist and a generalist competing with each other in an environment composed of two different habitats. The analysis shows that dispersal due to directed movement toward a favorable habitat and density dependent random movement both facilitate coexistence.  相似文献   

14.
The t-lethal haplotypes (t) found in house mouse (Mus musculus) populations are recessive lethals favored by gametic selection whereby male heterozygotes exhibit a non-Mendelian transmission ratio of about 95% t. The expected equilibrium frequency is 0.385; however, empirical values are lower, averaging close to 0.13. We examined the hypothesis that interdemic selection is the cause of the low empirical values by using a deme-structured simulation model that included overlapping generations, a realistic breeding system, differential deme productivity, and a large total population. We found that under some conditions interdemic selection could lower t frequency below 0.13 in the face of immigration rates up to 5%. Low frequencies were correlated with effective deme size (ne), regardless of whether ne was changed through changing deme size (n) or through changing the proportion of breeding adults. Earlier workers showed how the first two phases of interdemic selection (random genetic differentiation and mass selection) interacted to reduce the haplotype frequency, but here we show the importance of the third phase (differential productivity of demes) once demes are linked by dispersal. The effect of this phase is not due to the (negative) covariation between deme productivity and haplotype frequency, but occurs when differential deme productivity generates a difference in t frequency between the population of juveniles recruited into their natal deme and the population of juvenile dispersers. This difference was maximized when the average productivity of demes was low, either because few adult females bred at any one time and/or because fecundity was low. Contrary to an earlier prediction, male-biased dispersal also reduced haplotype frequency, and this probably stems from the relative excess of wild-type genotypes among dispersers compared to the deme residents. Another unexpected finding was that the randomly generated excess of heterozygotes (FIS < 0) found in small demes favored t haplotypes; however, the effect was only seen when the more powerful influence of the third phase of interdemic selection was removed. Simulations of neutral polymorphisms showed that a deme structure giving FST ≤ 0.6 is inconsistent with a haplotype frequency below 0.13. Based on current empirical estimates of FST (about 0.2), we concluded that immigration rates in the field are too high for interdemic selection alone to cause the observed deficit of lethal haplotypes. One factor that could combine with population structure effects is the observation that the transmission ratio is lowered to around 0.6 in litters produced from postpartum estrus (PPE). Incorporating this factor, we showed that interdemic selection could be effective in lowering the frequency of t below 0.13 when FST was above 0.43 even when migration rates were up to 10%. These results suggest that if empirical haplotype and FST estimates are accurate, then additional factors such as a lowered fitness of heterozygotes may be involved.  相似文献   

15.
Using the model of exploitative competition of R. H. MacArthur and R. Levins (1967, Amer. Natur. 101, 377–385), evolution at a gene locus which influences the niche position is considered. The locus has multiple alleles, and the contributions of the alleles to the genotypic value are additive. The resource spectrum and the utilization functions of the genotypes are assumed to be Gaussian. Evolution will make the mode of the niche converge to the resource optimum, as long as the allele contributions are small compared to the distance between the mode of the niche and the resource optimum. When this distance is of the same order of magnitude as the allele contributions, then the globally stable equilibrium will maintain at most two alleles in the population, unless the allele contributions are large. Classical overdominance is not needed to maintain polymorphism. This result predicts high linkage disequilibrium in similar multilocus models. It is concluded that intraspecific competition can be a powerful force in maintaining two-allele polymorphisms, and that it can maintain high linkage disequilibrium among closely linked loci.  相似文献   

16.
On the basis of systematic studies on the structure–activity relationships in arylpiperazine group of serotonin ligands, 12 new derivatives containing quinazolidin-4(3H)-one (14), 2-phenyl-2,3-dihydrophthalazine-1,4-dione (58) or 1-phenyl-1,2-dihydropyridazine-3,6-dione (912) fragments were synthesized. The majority of the tested compounds (2, 4, 7, 8 and 1012) showed a high affinity for 5-HT1A receptors (Ki=11–54 nM) and two (1, 2) were found active at 5-HT2A sites (16 and 68 nM, respectively). All the new 5-HT1A ligands tested in vivo revealed an antagonistic activity at postsynaptic 5-HT1A receptors, and three of them behaved as agonists at presynaptic ones. Additionally, both the meta-chlorophenylpiperazine derivatives containing quinazolidin-4-one fragment showed features of 5-HT2A receptor antagonists. The dual 5-HT1A/5-HT2A receptor ligand (2) was further tested for its potential psychotropic activity. It showed a distinct anxiolytic-like activity in a conflict drinking test in rats and the observed effect was more potent in terms of the active dose, than that produced by diazepam (used as a reference drug).  相似文献   

17.
4-(Substituted-benzylidine)-2-substituted-5,6-dihydrobenzo[h]quinazoline (5ap) and 4-(substituted-benzylidine)-2-substituted-3, 4, 5, 6-tetrahydrobenzo[h]quinazoline (6ap) have been synthesized from 2-(substituted-benzylidine)tetralone-1(3ad) and several substituted guanidine sulfates(4ad).These compounds were tested for their in vitro antileishmanial activity. The compounds 6i, 6f, 6g show promising antileishmanial activity against Leishmania donovani.  相似文献   

18.
R plasmids of a new incompatibility group, IncHII, determined the constitutive production of H pili, had high molecular weights, and determined tellurite resistance. They were designated IncHII because, during incompatibility tests, they sometimes eliminated or were eliminated by, previously described IncH plasmids, which they resembled in several respects. Nevertheless, stable and separate coexistence, i.e., compatibility, with plasmids of IncH1, IncH2, and IncH3 was demonstrated. The latter subgroups, members of which are all incompatible with one another, were distinguished on the basis of DNA-DNA hybridization experiments (N. D. F. Grindley, G. O. Humphreys, and E. S. Anderson, 1973, J. Bacteriol. 115, 387–398; A. F. Roussel, and Y. A. Chabbert, 1978, J. Gen. Microbiol. 104, 269–276.); it is proposed that they be called IncHI, the subgroups being HI1, HI2, and HI3.  相似文献   

19.
Resource availability influences sexual selection within populations and determines whether behaviours such as territoriality or resource sharing are adaptive. In Thoropa taophora, a frog endemic to the Atlantic Coastal Rainforest of Brazil, males compete for and defend limited breeding sites while females often share breeding sites with other females; however, sharing breeding sites may involve costs due to cannibalism by conspecific tadpoles. We studied a breeding population of T. taophora to determine (i) whether this species exhibits polygynous mating involving female choice for territorial males and limited breeding resources; (ii) whether limited breeding resources create the potential for male–male cooperation in defence of neighbouring territories; and (iii) whether females sharing breeding sites exhibit kin‐biased breeding site choice, possibly driven by fitness losses due to cannibalism among offspring of females sharing sites. We used microsatellites to reconstruct parentage and quantify relatedness at eight breeding sites in our focal population, where these sites are scarce, and in a second population, where sites are abundant. We found that at localities where the appropriate sites for reproduction are spatially limited, the mating system for this species is polygynous, with typically two females sharing a breeding site with a male. We also found that females exhibit negative kin‐bias in their choice of breeding sites, potentially to maximize their inclusive fitness by avoiding tadpole cannibalism of highly related kin. Our results indicate that male territorial defence and female site sharing are likely important components of this mating system, and we propose that kinship‐dependent avoidance in mating strategies may be more general than previously realized.  相似文献   

20.
Shifts between adaptive peaks, caused by sampling drift, are involved in both speciation and adaptation via Wright's “shifting balance.” We use techniques from statistical mechanics to calculate the rate of such transitions for a population in a single panmictic deme and for a population which is continuously distributed over one- and two-dimensional regions. This calculation applies in the limit where transitions are rare. Our results indicate that stochastic divergence is feasible despite free gene flow, provided that neighbourhood size is low enough. In two dimensions, the rate of transition depends primarily on neighbourhood size N and only weakly on selection pressure (≈sk exp(− cN)), where k is a number determined by the local population structure, in contrast with the exponential dependence on selection pressure in one dimension (≈exp(− cNs)) or in a single deme (≈exp(− cNs)). Our calculations agree with simulations of a single deme and a one-dimensional population.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号