To save content items to your account,
please confirm that you agree to abide by our usage policies.
If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account.
Find out more about saving content to .
To save content items to your Kindle, first ensure no-reply@cambridge.org
is added to your Approved Personal Document E-mail List under your Personal Document Settings
on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part
of your Kindle email address below.
Find out more about saving to your Kindle.
Note you can select to save to either the @free.kindle.com or @kindle.com variations.
‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi.
‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.
Genetic and cytological approaches have yielded significant insight into the mapping and organization of genes located in the heterochromatin of Drosophila melanogaster. To date, only a few of these genes have been molecularly characterized in detail, and their function unveiled. As a further step towards the identification of heterochromatic gene functions, we have carried out a cytological analysis of mitotic and meiotic cell divisions in mutants carrying different allelic combinations of l(2)41Aa, a gene located in the proximal heterochromatin of chromosome 2. Our results showed that larval brains of l(2)41Aa mutants display a high frequency of cells with irregularly condensed chromosomes. In addition, defective chromosome condensation was detected in male meiosis, consequently affecting chromosome segregation and giving rise to irregular spermatids. Taken together, these findings indicate that l(2)41Aa is a novel cell cycle gene required for proper chromosome condensation in both somatic and germ line cells.
We investigated the phenotypic plasticity of sternopleural bristle (SB) number as a function of growth temperature in isofemale lines from temperate (France) and tropical (Congo) populations of Drosophila melanogaster. We found concave reaction norms with a maximum in the middle of the thermal range, except in four African lines which exhibited a regularly decreasing response curve. Genetic variability (intraclass correlation) and evolvability (genetic CV, coefficient of variation) were independent properties and did not change with temperature. Residual, within-line variability was, however, strongly influenced by growth temperature, showing a U-shaped response curve and a minimum CV of 9% at 21·5 °C. As expected from a previously known latitudinal cline, maximum values (MV) were higher in temperate than in tropical flies. The temperature of maximum value (TMV) was observed at a higher temperature in the tropical population, in agreement with similar adaptive trends already observed for other quantitative traits. Significant negative correlations within each population were observed between a plasticity curvature parameter and MV or TMV. No difference in curvature was, however, observed between populations, in spite of their very different MVs.
Growth trajectories are a biological process important to plant and animal breeding, and to evolutionary genetic studies. In this article, we report the detection of quantitative trait loci (QTLs) responsible for growth trajectories in poplars that are used as a model system for the study of forest biology. These QTLs were localized on a genetic linkage map of polymorphic markers using a statistical mapping method incorporating growth-curve models. The effects of the QTLs on growth are described as a function of age, so that age-specific changes in QTL effects can be readily projected throughout the entire growth process. The QTLs identified display increased effects on growth when trees age, yet the timing of QTL activation is earlier for stem height than diameter, which is consistent with the ecological viewpoint of canopy competition. The implications of the results for breeding and silviculture are discussed.
A series of multivariate mixed-inheritance models is fitted to the data from an outbred-line pig cross commercially used in Norway. Each model accommodates information on polygenic (co)variances between F2 individuals and their F1 parents across the five traits through incorporation of a random animal effect. Considered traits relate to meat quality and are chosen following up the results from a previous evaluation, in which a putative quantitative trait locus (QTL) was identified on chromosome six that affects the amount of intramuscular fat (IMF), meat percentage, meat tenderness and smell intensity (Grindflek et al., 2001). An additional trait included in the model, based on results of other studies, is the backfat thickness. The analysed material comprises data scored for 305 F2 individuals, whereas marker information is available for F1 and F2 generations. Based on the results of the multivariate analysis with the mixed-inheritance model, it was possible to conclude that the evidence for QTLs for meat percentage, meat tenderness and smell intensity in the study of Grindflek et al. (2001) do not represent separate QTLs, but is caused by the fact that the applied pre-adjustment of trait values for polygenic effects failed properly to remove the polygenic variation. The QTL effect on IMF on chromosome six was confirmed.
We have isolated and characterized 77 novel microsatellites from two species, Drosophila dunni and Drosophila nigrodunni, which are closely related Caribbean-island endemics from the Drosophila cardini species group. These species are very distantly related to all other Drosophila from which microsatellites have previously been characterized. We find that the average length of microsatellites isolated in these species is quite small, with an overall mean length of 9·8 repeat units for dinucleotide microsatellites in the two study species. The nucleotide composition of dinucleotides differs between the two species: D. nigrodunni has a predominance of (AC/GT)n repeats, whereas D. dunni has equal numbers of (AC/GT)n and (AG/CT)n repeats. Tri- and tetranucleotide repeats are not abundant in either species. We assayed the variability of eight microsatellites in a closely related third species, Drosophila arawakana, using wild-caught individuals from the island of Guadeloupe. We found the microsatellites to be extremely variable in this population, with observed heterozygosities ranging from 0·541 to 0·889. DNA amplification trials suggest that these eight microsatellites are widely conserved across the D. cardini group, with five of the eight producing amplification products in every species tested. However, the loci are very poorly conserved over greater phylogenetic distances. DNA amplification of the microsatellite loci was unreliable in members of the closely related Drosophila quinaria, Drosophila calloptera, Drosophila guarani and Drosophila tripunctata species groups. Furthermore, these microsatellites could not be detected in the genome of Drosophila melanogaster, despite the conservation of microsatellite flanking regions at some loci. These data indicate that Drosophila microsatellite loci are quite short lived over evolutionary timescales relative to many other taxa.
Transposable elements (TEs) are genomic parasites that amplify their own representation on hosts' chromosomes by inserting into new positions. It is traditionally thought that their copy number is regulated by purifying selection that eliminates hosts with higher than average TE abundance. Here, we stress that selection due to beneficial or harmful interactions between TEs introduces a whole new dimension, with implications for TE evolutionary trajectories and TE loads on hosts. This framework poses new questions requiring conceptual and experimental advances. Considering primarily Drosophila data, we make a case for within host selection on TEs by thinking expansively about the lifecycle of several TE families.
‘Evolution Canyon’ on Mount Carmel, Israel, displays highly contrasting physical and biotic environments on a micro-geographic scale, and is a natural laboratory for investigating genetic responses to variable and extreme environments across species. Samples of Drosophila melanogaster and D. simulans were collected from three sites each on the north- and south-facing slopes of the canyon along altitudinal transects, and one site on the valley floor. Numbers of abdominal and sternopleural sensory bristles were recorded for each of these subpopulations in three thermal environments. In D. simulans, sternopleural bristle number exhibited micro-geographic differentiation between the north- and south-facing slopes, while abdominal bristle number was stable across subpopulations. In D. melanogaster, the magnitudes of the difference in mean sternopleural bristle number between the north- and south-facing slopes and of mean abdominal bristle number along the altitudinal gradients were both conditional on rearing temperature. Thus, the pattern of genetic variation between sites was consistent with underlying heterogeneity of genetic mechanisms for response to the same environmental gradients between traits and sibling species. In contrast, the genetic architecture of bristle number at the level of variation within populations was very similar between species for the same bristle trait, although the two traits differed in the relative contribution of genotype by temperature and genotype by sex interaction.
Positive and negative selection on indel variation may explain the correlation between intron length and recombination levels in natural populations of Drosophila. A nucleotide sequence analysis of the 3·5 kilobase sequence of the alcohol dehydrogenase (Adh) region from 139 Drosophila pseudoobscura strains and one D. miranda strain was used to determine whether positive or negative selection acts on indel variation in a gene that experiences high levels of recombination. A total of 30 deletion and 36 insertion polymorphisms were segregating within D. pseudoobscura populations and no indels were fixed between D. pseudoobscura and its two sibling species D. miranda and D. persimilis. The ratio of Tajima's D to its theoretical minimum value (Dmin) was proposed as a metric to assess the heterogeneity in D among D. pseudoobscura loci when the number of segregating sites differs among loci. The magnitude of the D/Dmin ratio was found to increase as the rate of population expansion increases, allowing one to assess which loci have an excess of rare variants due to population expansion versus purifying selection. D. pseudoobscura populations appear to have had modest increases in size accounting for some of the observed excess of rare variants. The D/Dmin ratio rejected a neutral model for deletion polymorphisms. Linkage disequilibrium among pairs of indels was greater than between pairs of segregating nucleotides. These results suggest that purifying selection removes deletion variation from intron sequences, but not insertion polymorphisms. Genome rearrangement and size-dependent intron evolution are proposed as mechanisms that limit runaway intron expansion.
The properties of alleles at quantitative trait loci (QTLs) contributing to variation in lifespan should be described to determine the mechanisms of evolution of life length and to predict its future changes. Previously, we and others conducted genome-wide screens for QTLs that segregate among one panel of recombinant inbred lines (RILs) using a dense molecular marker map. In non-stressful conditions, QTLs effecting the lifespans of virgin females and males were frequently sex specific. In an unrelated panel of RILs, the effects of QTLs in flies maintained in cages with mixed sexes were similar in both sexes. Here, we re-measured the lifespans of the former panel of RILs in cages with mixed sex cohorts. Lifespan declined owing to mating. The amount of decline correlated between sexes within lines. QTLs mapping to the intervals 15A–19C, 50B–57C, 63A–65A, and 96F–99B had similar effects on the lifespans of both males and females. These QTLs have previously been detected in virgin flies surveys and had sex- and/or environment-specific effects.
Gynodioecy, a genetic dimorphism of females and hermaphrodites, is pertinent to an understanding of the evolution of plant gender, mating and genetic variability. Classical models of nuclear gynodioecy attribute the maintenance of the dimorphism to frequency-dependent selection in which the female phenotype has a fitness advantage at low frequency owing to a doubled ovule fertility. Here, I analyse explicit genetic models of nuclear gynodioecy that expand on previous work by allowing partial male sterility in combination with either fixed or dynamically evolving mutational inbreeding depression. These models demonstrate that partial male sterility causes fitness underdominance at the mating locus, which can prevent the spread of females. However, if partial male sterility is compensated by a change in selfing rate, overdominance at the mating locus can cause the spread of females. Overdominance at introduction of the male sterility allele can be caused by high inbreeding depression and a lower selfing rate in the heterozygote, by purging of mutations by a higher selfing rate in the heterozygote, and by low inbreeding depression and a higher selfing rate in the heterozygote. These processes might be of general importance in the maintenance of mating polymorphisms in plants.
Finite polygenic models (FPM) might be an alternative to the infinitesimal model (TIM) for the genetic evaluation of pedigreed multiple-generation populations for multiple quantitative traits. I present a general flexible Bayesian method that includes the number of genes in the FPM as an additional random variable. Markov-chain Monte Carlo techniques such as Gibbs sampling and the reversible jump sampler are used for implementation. Sampling of genotypes of all genes in the FPM is done via the use of segregation indicators. A broad range of FPM models, some combined with TIM, are empirically tested for the estimation of variance components and the number of genes in the FPM. Four simulation scenarios were studied, including genetic models with 5 or 50 additive independent diallelic genes affecting the traits, and random selection or selection on one of the traits was performed. The results in this study were based on ten replicates per simulation scenario. In the case of random selection, uniform priors on additive gene effects led to posterior mean estimates of genetic variance that were positively correlated with the number of genes fitted in the FPM. In the case of trait selection, assuming normal priors on gene effects also led to genetic variance estimates for the selected trait that were negatively correlated with the number of genes in the FPM. This negative correlation was not observed for the unselected trait. Treating the number of genes in the FPM as random revealed a positive correlation between prior and posterior mean estimates of this number, but the prior hardly affected the posterior estimates of genetic variance. Posterior inferences about the number of genes should be considered to be indicative where trait selection seems to improve the power of distinguishing between TIM and FPM. Based on the results of this study, I suggest not replacing TIM by the FPM, but combining TIM and FPM with the number of genes treated as random, to facilitate a highly flexible and thereby robust method for variance component estimation in pedigreed populations. Further study is required to explore the full potential of these models under different genetic model assumptions.
Accurate and rapid methods for the detection of quantitative trait loci (QTLs) and evaluation of consequent allelic effects are required to implement marker-assisted selection in outbred populations. In this study, we present a simple deterministic method for estimating identity-by-descent (IBD) coefficients in full- and half-sib families that can be used for the detection of QTLs via a variance-component approach. In a simulated dataset, IBD coefficients among sibs estimated by the simple deterministic and Markov chain Monte Carlo (MCMC) methods with three or four alleles at each marker locus exhibited a correlation of greater than 0·99. This high correlation was also found in QTL analyses of data from an outbred pig population. Variance component analysis used both the simple deterministic and MCMC methods to estimate IBD coefficients. Both procedures detected a QTL at the same position and gave similar test statistics and heritabilities. The MCMC method, however, required much longer computation than the simple method. The conversion of estimated QTL genotypic effects into allelic effects for use in marker-assisted selection is also demonstrated.
Founder-origin probability methods are used to trace specific chromosomal segments in individual offspring. A haplotypic method was developed for calculating founder-origin probabilities in three-generation outbred pedigrees suited to quantitative trait locus (QTL) analysis. Estimators for expected founder-origin proportions were derived for a linkage group segment, an entire linkage group and a complete haplotype. If the founders are truly outbred, the haplotypic method gives a close approximation when compared with the Haley et al. (1994) method that simultaneously uses all marker information for QTL analysis, and it is less computationally demanding. The chief limitation of the haplotypic method is that some information in two-allele intercross marker-type configurations is ignored. Informativeness of marker arrays is discussed in the framework of founder-origin probabilities and proportions. The haplotypic method can be extended to more complex pedigrees with additional generations.
Naturally occurring genetic variation was quantified for survival time of adult Drosophila melanogaster exposed to chronic ingestion of the drugs nicotine, caffeine, dopamine, tyramine and octopamine. Responses to nicotine, tyramine and octopamine were genetically correlated in both sexes, whereas caffeine response correlated with starvation resistance. However, there is also genetic variation that is specific for each of the drugs. Females tended to be more resistant than males to nicotine and caffeine but sex-by-genotype interactions were also seen for these drugs and for the response to dopamine. An unusual and complex genetic architecture was observed in crosses between lines with different responses to caffeine ingestion. Additive and dominance components were clearly seen from the analysis of F1 individuals, but increased female resistance to caffeine in backcross generations and increased male sensitivity in F2 generations confused the interpretation of possible epistatic contributions.
The aim of this paper is to investigate the effect of deleterious mutations in a hybrid zone maintained by selection against hybrids. In such zones, linkage disequilibria among hybrid depression loci, resulting from a balance between migration and selection, are crucial in maintaining the barrier because they allow each locus, in addition to its own selection coefficient, to cumulate indirect selective effects from other loci. Deleterious alleles produce heterosis and increase by this means the effective migration rate in structured populations. In a hybrid zone, they therefore contribute to decrease linkage disequilibria as well as the barrier to gene flow imposed by hybrid depression. However, deleterious mutations have no effect: (i) when selection against hybrids is weak, because linkage disequilibria are small even without heterosis in this case, or (ii) when selection against hybrids is so strong that it overwhelms heterosis. On the other hand, with moderate selection against hybrids, the decrease in the strength of the barrier due to heterosis may reach detectable levels, although it requires relatively small population sizes and/or migration rates. The effect is expected to be small and only within small genomes where loci are tightly linked can it become strong. Nevertheless, neglecting mutational load may to some extent obscure the estimations of selective parameters based either on artificial F1 crosses or on cline characteristics.
PRIMER When we dissect chromatin, we arrive at the highest resolution, which is the DNA wrapped around histones. The structure of DNA, the double helix, needs no elaborate introduction. In this chapter, the basic elements that contribute to the 3-D structure of DNA are presented. The helical DNA can, in fact, assume different structural forms, and we discuss them along with their significance. However, not all DNA exists as double-stranded. At the ends of the chromosomes, DNA is single-stranded and is characterized by a specific nucleotide sequence. The ends of the chromosomes are called telomeres, and they can assume a distinct 3-D structure. This structure is presented along with the structure of proteins that binds telomeric sequence and plays roles in their stability that eventually affect the length of chromosomal DNA and its role in replication and cell division.
THE THREE-DIMENSIONAL STRUCTURE OF DNA
The 3-D structure of DNA is undoubtedly the most recognizable structure of a biological molecule to scientists and nonscientists alike. The determination of the 3-D structure of DNA in 1953 stunned the scientific world and completed the race for understanding how genetic information is passed on to progeny. The DNA is a simple periodic spiral structure made up of two helices. Each helix is made up of sequence of four bases connected via phosphate bonds, and the two strands are held together by a specific interaction between the bases.
PRIMER Transcription in eukaryotes is more complicated than in prokaryotes. In eukaryotes, there must be tissue-specific gene transcription because many different cell types need to be created. Therefore, transcription must be selective in different cells and tissues. This selectivity is possible by the cell- or tissue-specific transcriptional factors. These factors can bind sequences present in promoters or enhancers of genes that will be specifically expressed in a particular cell type. The basal transcriptional machinery, which is assembled at the promoters of genes and is responsible for the recruitment of RNA polymerase, is also more complex in eukaryotes. In addition, in eukaryotes there are three RNA polymerases, each one transcribing a particular set of genes.
As in Chapter 5, I begin this chapter by presenting the 3-D structure of eukaryotic RNA polymerase II, which transcribes the genes that produce mRNAs. A comparison is made with the prokaryotic RNA polymerase. I then describe the factors that constitute the basal transcriptional machinery, their 3-D structure, and their interaction with DNA, RNA polymerase II, and each other. The reader should be able to obtain a very informative picture of this assembly that leads to the initiation of eukaryotic transcription. I then proceed to the regulation of transcription. As I mentioned earlier, regulation of transcription is unique for each of the more than 200 different cell types. Covering everything we know about regulation is, therefore, overwhelming.
PRIMER DNA replication involves many proteins acting at different steps. DNA polymerases are the main enzymes needed to create new strands from the parental ones. However, DNA polymerase alone cannot complete the whole task of replication. When the double-stranded DNA melts at the origin of replication to expose the single strands to the polymerizing enzyme, the fork must continue to unwind until the termination signals have been reached. This part in replication is performed by helicases, and the topology ahead of the fork (mainly coiling and superwinding of the DNA) is controlled by topoisomerases. Finally, termination is achieved by the interaction of both helicases and termination proteins. Editing is another important function during replication to ensure that misincorporations are eliminated and, thus, mutations are avoided. Also, other proteins are involved in securing the processivity of DNA polymerase and the processing of the Okazaki fragments. The 3-D structure of all the major players participating in the entire process of replication is for the most part known. For some of them, we also know how they interact with each other or with DNA. This enables us to visualize replication in a 3-D journey. This chapter introduces the 3-D structure of DNA polymerases, the mechanism of incorporation of the incoming nucleotide, and how DNA polymerase interacts with DNA and other proteins to achieve processivity and editing. We then examine the structure of different helicases, their interaction with DNA, and the mechanism of unwinding as suggested by the 3-D structures.