Download Software Yvert Et Tellier 2011 Pdf

Download Software Yvert Et Tellier 2011 Pdf Average ratng: 6,6/10 6777 reviews

Create a book Download as PDF Printable version. Free download here yvert et tellier stamp catalogues manual yvert pdf ebook. Yvertmanual service lancer evo 3jee online key answer of 12 4. Yvert stamp catalogue, yvert stamp catalogue download Yvert Stamp Catalogue Pdf Download - http://urllio.com/ycvrt.

CEGMA and BUSCO results for the Berthelot's pipit genome assembly.Table S2. Nucleotide diversity across groups of Berthelot's pipits and tawny pipits.Table S3. Outlier SNPs with PIP 0.1 in BSLMM analyses for bill length and malaria in the 'Berthelot's' datasetTable S4. Outlier SNPs from Eigenvector 2 of EigenGWAS analysis.Figure S1. Nucleotide diversity in Tawny pipits and Berthelot's pipit archipelagosFigure S2.

Cross‐validation (CV) error for K = 1 to K = 13 clusters, calculated by Admixture analysis of Berthelot's pipit populationsFigure S3. Admixture analysis at K = 7 to K = 13 clusters for Berthelot's pipitFigure S4. Pairwise genetic distance in relation to geographical distance across Berthelot's pipit populations in the Canary IslandsFigure S5. The relationship between BSLMM outlier SNP pairwise F ST and A) pairwise genome‐wide F ST; and B) pairwise geographic distance between all pairs of Berthelot's pipit populationsFigure S6. Genetic associations with tarsus length.Figure S7. Genetic associations with malaria and bill length in the 'Berthelot's' datasetFigure S8. Selection across archipelagos.Figure S9.

Region of interest for bill length on chromosome 5Figure S10. Individual observed heterozygosity averaged across SNPs taken from EigenGWAS outliers (candidate regions ‐ see main text) versus the rest of the genome (non‐candidate regions)Figure S11. Linkage disequilibrium between all pairs of outlier SNPs from EigenGWAS Eigenvector 1 and malaria BSLMM analysis.

Island species provide excellent models for investigating how selection and drift operate in wild populations, and for determining how these processes act to influence local adaptation and speciation. Here, we examine the role of selection and drift in shaping genomic and phenotypic variation across recently separated populations of Berthelot's pipit ( Anthus berthelotii), a passerine bird endemic to three archipelagos in the Atlantic. We first characterized genetic diversity and population structuring that supported previous inferences of a history of recent colonizations and bottlenecks. We then tested for regions of the genome associated with the ecologically important traits of bill length and malaria infection, both of which vary substantially across populations in this species. We identified a SNP associated with variation in bill length among individuals, islands, and archipelagos; patterns of variation at this SNP suggest that both phenotypic and genotypic variation in bill length is largely shaped by founder effects.

Malaria was associated with SNPs near/within genes involved in the immune response, but this relationship was not consistent among archipelagos, supporting the view that disease resistance is complex and rapidly evolving. Although we found little evidence for divergent selection at candidate loci for bill length and malaria resistance, genome scan analyses pointed to several genes related to immunity and metabolism as having important roles in divergence and adaptation. Our findings highlight the utility and challenges involved with combining association mapping and population genetic analysis in nonequilibrium populations, to disentangle the effects of drift and selection on shaping genotypes and phenotypes. Sampling locations, genetic diversity, and linkage disequilibrium in Berthelot's and tawny pipits.

(A) Map of Berthelot's pipit and tawny pipit sampling locations. (B) The relationship between linkage disequilibrium and base‐pair distance for SNPs in the Canary Islands, Madeira, and Selvagens archipelagos.

The red line shows a general additive model, with the shaded band indicating 95% confidence intervals.We chose 20 samples per population (22 from the lowland Tenerife population) for RAD‐seq, with efforts made to reduce the probability of sampling relatives by maximizing geographical coverage within each island, and to equalize the number of males and females where possible. In addition, 16 samples from tawny pipits were included; these were collected in Spain ( n = 11), Mauritania ( n = 4), and Morocco ( n = 1; Fig.

A) following the same catching, measuring, and sampling protocols as Berthelot's pipit. Sequencing and bioinformatics were performed using the ddRAD protocol by DaCosta and Sorenson , with loci mapped to the zebra finch genome ( Taeniopygia guttata, version 3.2.4; Warren et al. Where initial mapping was unsuccessful, a draft Berthelot's pipit genome was used to improve mapping to the zebra finch genome (see Supplementary Methods). Following demultiplexing (Supplementary Methods), samples were grouped into five datasets, as outlined in Table. These were created to maximize the number of loci at different levels of clustering within our dataset (Supplementary Methods; Table ). DatasetAll pipitsBerthelot'sBerthelot's HWECanary IslandsMadeiraSelvagensn Samples826020n Loci1918Read depth87195176Density of loci in zebra finch genome (%)0.0160.0250.0240.0300.0270.017Sample locationsCanary IslandsXXXXMadeiraXXXXSelvagensXXXXTawny pipitsXFiltering stepsFiltering AXFiltering BXXXXXFiltering CXXXXXFiltering DXAnalysesNucleotide diversityXLinkage disequilibriumXXXAdmixtureXBSLMM/LMMXXEigenGWASX. POPULATION GENETICS ANALYSISWe first compared the overall levels of genetic diversity among the different Berthelot's pipit archipelagos and the tawny pipit, using the “All Pipits” dataset.

We used the approach outlined by Lozier to calculate per‐SNP nucleotide diversity (π SNP) at 1826 loci present in all samples, and the overall nucleotide diversity per locus (π RAD) as the sum of the nucleotide diversity at every site along a RAD locus, including invariable sites, divided by the length of the locus (excluding loci. GENETIC ASSOCIATIONS WITH TRAITSWe used two genome‐wide association study (GWAS) analyses to identify loci associated with bill length and malaria infection. A central issue with GWAS is accounting for population structure (Marchini et al.

) and we used two approaches to deal with this. First, we restricted association analyses to the “Canary Islands” dataset, in which population structure is limited (see Results). Secondly, we used kinship/relatedness matrices to statistically control for any population structure among the Canary Islands. We first used a univariate linear‐mixed model (LMM) implemented in the software package GEMMA (Zhou and Stephens ). This model assumes that every genetic variant affects the phenotype, which would suit a highly polygenic genetic architecture, and accounts for population structure with a centered kinship matrix calculated in GEMMA.In our second analysis, we performed association mapping with a Bayesian sparse linear‐mixed model (BSLMM; Zhou et al. ), again using GEMMA and the “Canary Islands” dataset.

BSLMM combines linear‐mixed models with sparse regression models, giving the benefits of each when the underlying genetic architecture of the trait (many genes of small effect vs few genes of large effect) is unknown (Zhou et al. Population structure is controlled for with a relatedness matrix as a covariate of the model. The output includes estimates of the hyperparameters PVE (the proportion of the phenotypic variance explained), PGE (the proportion of PVE that can be explained by SNPs with a nonzero effect on phenotypic variation), and the number of SNPs that explain the PVE. Each SNP is assigned a posterior inclusion probability (PIP), which is the proportion of times the SNP has a nonzero effect on phenotypic variation. We used a threshold of PIP 0.1 to identify candidate SNPs associated with phenotypes (Chaves et al. Each analysis was run for 20 million iterations with a burn‐in of 5 million.

This was repeated 10 times, and the results were averaged across runs. We subsequently repeated the BSLMM analysis using the “Berthelot's” dataset, to discover whether the increase in power gained from a larger sample size improved the ability to detect loci associated with either bill length or malaria infection.To check whether SNPs associated with bill length also covary with overall body size, we also ran a BSLMM with tarsus length as a candidate trait. Tarsus length is the strongest predictor of overall body size that can be accurately measured from a live bird (Senar and Pascual ).

Bill and tarsus length were only weakly correlated in our dataset ( r = 0.13, P = 0.04).To identify genes located near outlier SNPs, we downloaded Ensembl gene predictions for zebra finch , and viewed regions of interest using the University of Santa Cruz Genome Browser.We examined how selection and drift shaped variation at candidate SNPs, as follows. For the most significant SNPs identified in the BSLMM analyses (PIP 0.1), we calculated allele frequencies for each island population and compared this against population‐level trait variation (mean bill length and malaria prevalence), using Pearson correlations. We also calculated pairwise F ST individually for BSLMM outliers, and compared outlier F ST and genome‐wide F ST, using partial Mantel tests. To better visualize how variation at outlier SNPs was partitioned among populations, we also tested for isolation‐by distance at outlier SNPs, using the same approach described above for genome‐wide F ST. EigenGWAS ANALYSISWe used EigenGWAS (Chen et al. ), implemented in the program GEAR , to separate loci under divergent selection from the genome‐wide effects of drift between the archipelagos, using the “Berthelot's” dataset.

This analysis uses eigenvector decomposition on the genotype data to characterize the underlying population genetic structuring, without the need to define discrete populations. This then enables the identification of SNPs under selection across gradients of population structure. EigenGWAS provides adjusted P values that use the genomic inflation factor λ GC (Devlin and Roeder ) to correct for population stratification (i.e., drift), to avoid interpreting ancestry‐informative markers as loci under selection. Genes in the zebra finch genome located near to outlier SNPs were identified as above.

We then calculated F ST in PLINK 1.9 (Chang et al. ), grouping samples according to eigenvector clustering, to examine the relationship between EigenGWAS significance and genetic divergence. Finally, we tested whether significant BSLMM and EigenGWAS SNPs were more closely located than expected by chance using randomization tests (Supplementary Methods). RAD LIBRARIES AND GENOME SUMMARYThe three RAD libraries produced 277,627,890 (first), 202,466,438 (second), and 184,660,891 (third) raw reads with GC contents of 52%, 51% and 50%, respectively. Following filtering, we produced six datasets containing between 1826 and 3330 polymorphic RAD loci per dataset, at a median read depth of between 176 and 198 reads per sample per locus (Table ).We created a draft Berthelot's pipit genome that we used to improve our ability to map RAD loci to the zebra finch genome.

We used DISCOVAR de novo (Weisenfeld et al. ) to produce an assembly 1,153,192,274 bp long (94.3% of the length of the zebra finch genome), with a contig N50 of 355,835 bp (see Supplementary Results and Table S1 for further details). GENETIC DIVERSITY AND POPULATION STRUCTUREWe used two measures of nucleotide diversity, π SNP and π RAD, to compare genetic diversity between the three separate Berthelot's pipit archipelagos and tawny pipits (Table S2, Fig. In line with expectations based on population sizes, tawny pipits had the highest levels of diversity (π SNP = 0.104, π RAD = 0.0049), followed by Berthelot's pipit populations in the Canary Islands (π SNP = 0.012, π RAD = 0.0005), Madeira (π SNP = 0.008, π RAD = 0.0004), and Selvagens (π SNP = 0.006, π RAD = 0.0003). There was moderate divergence between Deserta Grande and the rest of the Madeiran archipelago (see below), however, exclusion of Deserta Grande had little impact on measures of diversity (Table S2).This trend was reflected throughout Berthelot's pipit archipelagos for linkage disequilibrium (LD; Fig.

Baseline levels of LD were lowest in the Canary Islands and highest in Selvagens, with the rate of decay sharpest in the Canary Islands and shallowest in Selvagens.Population admixture analysis (Fig. ) first separated Madeira from the Canary Islands and Selvagens ( K = 2) and then into the three archipelagos ( K = 3). Further population structuring beyond the archipelago level was seen with Deserta Grande diverging from the other islands in the Madeiran archipelago ( K = 4), and an east–west gradient in admixture across the Canary Islands ( K = 5; Fig. The most likely number of clusters determined by cross‐validation error was four (Fig. S2), however similar cross‐validation errors were found for K = 3 (0.434), K = 4 (0.431), and K = 5 (0.434). ADMIXTURE plots for K = 7 to K = 13 are shown in Fig. We found a moderate signal of isolation by distance within the Canary Island archipelago (Mantel test, r = 0.37, P = 0.002; Fig.

S4), although levels of structure within the archipelago were generally low ( F ST. Admixture analysis at K = 2 to K = 6 clusters for Berthelot’s pipit. Vertical bars represent individual pipits, and are colored by their assignment to each one of K clusters. Solid black lines indicate separate populations; dashed black lines indicate separate archipelagos. Populations from left to right: Canary Islands – El Hierro (EH), La Palma (LP), Gomera (GOM), Teide (TEID), Tenerife (TF), Gran Canaria (GC), Fuerteventura (FV), Lanzarote (LZ), Graciosa (GRA); Selvagens – Selvagem Grande (SG); Madeira – Madeira (M), Porto Santo (PS), Deserta Grande (DG). ASSOCIATION STUDIESBoth LMMs and BSLMMs were used to find regions of the genome associated with malaria and bill length in the “Canary Islands” dataset, while accounting for population structure. The LMM analysis failed to find any SNPs significantly associated with either malaria or bill length when using a Bonferroni‐corrected significance threshold of P 0.1, after controlling for population structure (Fig.

Of these outlier SNPs, the top two also had the smallest P values in the LMM (Table ). The strongest association was found for a SNP on chromosome 10 (5239s1), approximately 2000 bp from the IL‐16 gene. The next strongest association was with a SNP on chromosome 20 (7259s1) situated within an intronic region of RIMS4.

Both of these SNPs had “A” and “T” alleles. The “AA” and “TT” homozygotes were associated with the highest and lowest levels of malaria, respectively, while heterozygotes had intermediate malaria prevalence (Fig. We then examined how population‐level allele frequencies at these SNPs were related to population‐level prevalence of malaria. Within the Canary Islands, there was a clear correlation between malaria prevalence and allele frequency for SNP 5239s1 ( r = –0.76, P = 0.02); however, this pattern was not observed when the populations from the bottlenecked northern archipelagos were included ( r = –0.35, P = 0.25; Fig. Despite showing an individual‐level association between malaria prevalence and genotype (Fig.

B), the population‐level allele frequency at SNP 7259s1 was not significantly correlated with malaria infection in either the Canary Islands ( r = –0.59; P = 0.09) or across all populations ( r = –0.50; P = 0.09; Fig. Estimates of F ST revealed that patterns of structure at both of these SNPs was correlated with genome‐wide F ST across all Berthelot's pipit populations (partial Mantel tests, 5239s1: r = 0.40, P = 0.007; 7259s1: r = 0.46; P = 0.01; Fig. Outlier F ST values were correlated with pairwise geographic distance (5239s1: r = 0.30, P = 0.009; 7259s1: r = 0.36; P = 0.01; Fig. Genetic associations with malaria infection.

(A) Manhattan plot of BSLMM analysis of malaria infection in the Canary Islands. Red line indicates the threshold of posterior inclusion probability (PIP) = 0.1. For the top two outlier SNPs: (B) the association between genotype and the proportion of individuals infected with malaria in the Canary Islands. Numbers above bars indicate the number of individuals with each genotype; and (C) per‐population minor allele frequency and the proportion of infected individuals, colored by archipelago (C = Canary Islands, M = Madeira, S = Selvagens). The BSLMM for bill length found that genotype explained a median of 67.9% of phenotypic variation (95% CI 14.0–99.7%), with 14.3% explained by SNPs with nonzero effects (although again the latter had wide credible intervals; 0.0–85.6%).

The strongest association with bill length was for a SNP on chromosome 5 (Fig. At this SNP (8914s1), individual bill length decreases with increasing number of copies of the “G” allele, and heterozygotes had intermediate bill length (Fig. This region was not associated with tarsus length (Fig.

S6A), and there were no SNPs that showed strong associations for both bill length and tarsus length (Fig. At the population level, mean bill length decreases with increasing frequency of the “G” allele at the most significant BSLMM SNP, both in the Canary Islands ( r = –0.69, P = 0.04), and across all Berthelot's pipit populations ( r = –0.82, P = 0.0006; Fig.

After controlling for archipelago‐level effects, patterns of F ST at this SNP were strongly correlated with genome‐wide F ST ( r = 0.64, P = 0.002; Fig. S5A), and with geographic distance ( r = 0.48, P = 0.002; Fig. Genetic associations with bill length and selection across archipelagos.

(A) Manhattan plot of BSLMM analysis of bill length in the Canary Islands. Red line indicates the threshold of posterior inclusion probability (PIP) = 0.1. For the top outlier SNP: (B) the association between genotype and bill length in the Canary Islands. Numbers above box plots indicate the number of individuals with each genotype; and (C) per‐population minor allele frequency and mean bill length, colored by archipelago (C = Canary Islands, M = Madeira, S = Selvagens). (D) Manhattan plot of EigenGWAS eigenvector 1.

Red line indicates the Bonferroni‐corrected P value threshold.Repeating the BSLMM analyses using the “Berthelot's” dataset resulted in higher PIPs for the outlier SNPs mentioned above, but also identified a number of additional outlier loci (Supplementary Results, Table S3, Fig. DIFFERENTIAL SELECTION ACROSS ARCHIPELAGOSWe used EigenGWAS to detect loci putatively under divergent selection between the archipelagos in the “Berthelot's” dataset.

The first eigenvector (EV1) separated the Madeiran archipelago from the Canary Islands and Selvagens, and the second (EV2) separated Selvagens from Madeira and the Canary Islands (Fig. The first two eigenvectors captured a large proportion of the genetic variation, with eigenvalues of 31.1 and 14.6, respectively. Both EV1 and EV2 had considerable genomic inflation factors (26.5 and 9.7, respectively), suggesting high levels of population structuring (a genomic inflation factor ≤ 1 indicates no population structuring; Hinrichs et al. ), so we used the adjusted P values calculated by EigenGWAS that use genomic control to account for population structure.Using a Bonferroni‐corrected P. None of the candidate SNPs identified in the BSLMM analyses was identified as outliers in the EigenGWAS analysis. An outlier SNP in the EigenGWAS analysis of EV1 was located close to the region on chromosome 5 that showed an association with bill length (Fig.

However, closer inspection of the region revealed that the most significant SNP was 2,932,270 bp from the most significant BSLMM SNP (Fig. S9), which was not closer than expected by chance (randomization test, P = 0.1). A second SNP in the region also had an elevated (but not significant) PIP value in the BSLMM analysis and was much closer to the significant EigenGWAS candidate SNP (444,885 bp; Fig. DiscussionUsing RAD sequencing, we examined fine‐scale population structure among Berthelot's pipit populations, finding within‐archipelago genetic structuring between island populations, as well as confirming the previously detected patterns of among‐archipelago variation. After controlling for demographic history, we identified candidate SNPs associated with the ecologically important traits of bill length and malaria resistance.

Examining allele frequency variation at candidate SNPs revealed that: (i) population‐level candidate SNP variation was related to population‐level trait variation for bill length both within the founder Canary Islands archipelago and throughout all archipelagos, but only within the Canary Islands for malaria; and, (ii) candidate SNP variation was correlated with genome‐wide variation. Finally, although we found little evidence for divergent selection on bill length or malaria resistance, we found signatures of divergent selection between the archipelagos across the genome, including at genes involved in metabolism and immune function.Comparisons of genome‐wide diversity in Berthelot's and tawny pipits provided support for previous inferences of colonization history from microsatellites and mitochondrial DNA (Illera et al.; Spurgin et al. The past colonization history and associated bottlenecks are reflected in linkage disequilibrium (LD) within the archipelagos, indicating a larger, more outbred population in the Canary Islands, compared with smaller, more inbred populations in Madeira and Selvagens. This is in concordance with previous estimations of effective population size using approximate Bayesian Computation, which predicted contemporary effective population sizes of ∼4000, ∼3000, and ∼400 individuals for the Canary Islands, Madeira, and Selvagens, respectively (Spurgin et al.

Using population admixture analysis, we were able to further describe population structure: Deserta Grande diverged from the rest of the Madeiran archipelago, possibly suggesting a further bottleneck on Deserta Grande. We also detected a weak east–west gradient in population structure within the Canary Islands, which was supported by the finding of isolation‐by‐distance throughout this archipelago.We next aimed to identify the genetic regions associated with bill length and malaria resistance—two traits of major evolutionary importance (Smith et al.; Daszak; Grant and Grant ). The high level of LD relative to outbred populations (e.g., Kardos et al. ) has enabled us to detect associations with these two polygenic traits in Berthelot's pipit, with a modest number of markers. The strongest association with bill length was found at a SNP in a region on chromosome 5. This SNP (8914s1) is not located within a gene, which is perhaps not surprising, given (i) the density of our marker set, and (ii) the fact that many SNPs associated with polygenic traits are located in regulatory regions (Maurano et al.

The wider region surrounding SNP 8914s1 and a significant EigenGWAS SNP (see below) contains, among others, the genes BMP4 and OTX2. BMP4 is a clear candidate for involvement in bill morphology, as studies in Darwin's finches, chickens, and ducks have experimentally demonstrated its role in beak development (Abzhanov et al.; Wu et al. OTX2 plays a crucial role in craniofacial development across jawed vertebrates (Kimura et al. ), and mutations in this gene and structural variants within the wider genomic region have been linked with craniofacial abnormalities in mice and humans (Hide et al.; Zielinski et al. While BMP4 and OTX2 are good candidates, we stress that further work is required to identify which specific genes affect variation in bill length.

Further, bill length is almost certainly a polygenic trait (e.g., Abzhanov et al.; Lamichhaney et al., ), and many causative loci have likely gone undetected in this study.There are now several examples linking variation at avian bills, at both the genetic and phenotypic levels, which have shown natural selection to be the primary driver of bill shape variation (Lamichhaney et al.; Bosse et al. Here, we revealed a striking relationship between genetic variation at our candidate bill‐length SNP 8914s1, and population‐level variation in bill length; the long‐bill allele identified using birds from the Canary Islands is at near fixation in the Madeiran Islands, where birds have the longest bills (Fig. We then found that variation at SNP 8914s1 was correlated with genome‐wide variation among populations (Fig. S5) with the lowest levels of diversity in the bottlenecked Madeiran and Selvagens populations (Fig. Combined with the finding that SNP 8914s1 was not identified in the EigenGWAS analysis, this suggests that founder effects are likely to have played a major role in shaping bill length variation among Berthelot's pipit populations. This is consistent with our previous phenotypic work on this species (Spurgin et al. ), but only now have we been able to show how founder effects simultaneously affect phenotypes and underlying genotypes—indeed, very few studies have done so.Despite a clear role for founder effects in shaping bill length variation, we did find that genetic structuring at SNP 8914s1 was higher than the genome‐wide average, particularly across the Canary Islands and Madeira archipelagos (Fig.

We also detected an EigenGWAS outlier SNP (8322s2) putatively under divergent selection within the wider BMP4/ OTX2 bill‐length candidate region (Fig. Furthermore, we might not expect to be able to identify genotype–phenotype associations at SNPs under strongest divergent selection, as these are expected to have low variability within populations (indeed this is the case in our data—Fig. S10) and detecting rare genetic variants associated with phenotypes is very difficult (Li and Leal ). Thus, while the majority of observed bill length variation among archipelagos can be explained by neutral forces, we cannot rule out that selection has occurred for longer bills in Madeira (or shorter bills in southerly populations). Further research is required to identify whether selection has played any role in shaping beak morphology, and the underlying genetic variants associated with this key ecological trait.Malaria can act as a strong selective force (Warner; Van Riper et al.; Ortego et al.; Knowles et al. ), and drive the evolution of increased tolerance to infection (Atkinson et al.

Genotypic associations with malaria infection in individual Berthelot's pipits revealed two outlier SNPs with high levels of significance (Table ). The first of these was on chromosome 10, approximately 2000 bp from IL‐16. This gene encodes a proinflammatory cytokine that has been implicated in susceptibility to various inflammatory disorders (Gao et al.; Wu et al. ), and interacts with other cytokines that are associated with malaria infection (Kern et al.; Mathy et al.; Lyke et al. The second outlier SNP was found on chromosome 20, within RIMS4.

Research on this gene is lacking, though it has been found to be overexpressed in breast cancer tumor cells (Abba et al. ).Examining allele frequency variation at the malaria resistance candidate SNPs revealed a complex set of relationships. Despite finding a clear relationship of malaria infection with both individual‐ and population‐level SNP variation in the Canary Islands, this pattern did not hold when the Madeiran and Selvagens archipelagos were included (Fig.

Further, our EigenGWAS analysis did not detect signatures of divergent selection between archipelagos in the regions surrounding the malaria candidates RIMS4 or IL‐16. In contrast to bill length, which seems more likely to have a conserved genetic architecture, resistance to malaria is probably a rapidly evolving trait, and candidate‐gene studies have shown that different alleles can be associated with malaria infection among different populations (Bonneaud et al. Variation at both malaria candidate SNPs was correlated with genome‐wide variation (Fig. S5) and likely to be shaped by founder effects. However, our study suggests that, because the genetic basis of malaria resistance is likely to vary among populations (as expected with rapid host–pathogen coevolution, Slade and McCallum ), there is unlikely to be a simple relationship between resistance to malaria infection and population demography in wild populations.The EigenGWAS analysis detected putative signatures of selection among the archipelagos at a number of locations around the genome (Table ). The significant EigenGWAS SNPs were all in high LD with one another (Fig.

S11), suggesting that these sites have evolved nonindependently. Also important to note is that many of these SNPs are not situated close to candidate genes. Some of these could situated close to features such as trans‐regulatory elements or structural variants (Yvert et al.; Manolio et al. ), or could be false positives that have arisen as a result of population bottlenecks (Akey et al. Sppngebob indo.mp4.

Our analyses suggest that the EigenGWAS analysis was more conservative than a simple F ST‐based approach to detecting selection, with numerous F ST outliers having nonsignificant EigenGWAS P values after employing a genomic control (Fig. This is in contrast to other studies that have shown almost perfect correlations between F ST and EigenGWAS P values (Chen et al.; Bosse et al.

), and suggests that EigenGWAS may be more conservative when levels of drift are high. However, it is unlikely that all false positives are accounted for, and further research is now needed to understand how EigenGWAS performs in bottlenecked populations.Confounding effects of bottlenecks aside, a substantial number of the SNPs identified in our EigenGWAS analyses are likely to be “true” positives, suggesting that traits other than bill length and malaria resistance may have been under selection among archipelagos.

We identified SNPs located within loci involved in immune function (e.g., PREX1, Welch et al.; MAP3K14, Thu and Richmond ), or with potential to enable adaptation to climate (e.g., DLAT, Blier and Guderley; SNX29, Sung et al. These findings are reflected in other studies that utilize genome scans to detect signatures of selection. Adaptation to toxic food sources in an isolated population of Drosophila yakuba was detected in a genome scan between island and mainland populations (Yassin et al.

Divergent selection for pathogen‐resistance candidate genes has been found in Daphnia (Bourgeois et al. ) and across a bank vole range expansion (White et al. ), and pathogens were the strongest selective pressure identified in human evolutionary history (Fumagalli et al.

Likewise, signatures of divergent selection have been detected for genes related to metabolic processes along latitudinal gradients in numerous species (Sezgin et al.; Hancock et al.; Pujolar et al. This suggests that there may be general patterns in the types of genes that show divergence between populations, with similar processes of adaptation across species. Future studies investigating links between genotypes under selection and ecologically important phenotypic traits will provide further support for the role of adaptation in driving patterns of biodiversity. CEGMA and BUSCO results for the Berthelot's pipit genome assembly.Table S2. Nucleotide diversity across groups of Berthelot's pipits and tawny pipits.Table S3. Outlier SNPs with PIP 0.1 in BSLMM analyses for bill length and malaria in the 'Berthelot's' datasetTable S4. Outlier SNPs from Eigenvector 2 of EigenGWAS analysis.Figure S1.

Nucleotide diversity in Tawny pipits and Berthelot's pipit archipelagosFigure S2. Cross‐validation (CV) error for K = 1 to K = 13 clusters, calculated by Admixture analysis of Berthelot's pipit populationsFigure S3. Admixture analysis at K = 7 to K = 13 clusters for Berthelot's pipitFigure S4.

Pairwise genetic distance in relation to geographical distance across Berthelot's pipit populations in the Canary IslandsFigure S5. The relationship between BSLMM outlier SNP pairwise F ST and A) pairwise genome‐wide F ST; and B) pairwise geographic distance between all pairs of Berthelot's pipit populationsFigure S6. Genetic associations with tarsus length.Figure S7.

Genetic associations with malaria and bill length in the 'Berthelot's' datasetFigure S8. Selection across archipelagos.Figure S9. Region of interest for bill length on chromosome 5Figure S10.

Individual observed heterozygosity averaged across SNPs taken from EigenGWAS outliers (candidate regions ‐ see main text) versus the rest of the genome (non‐candidate regions)Figure S11. Linkage disequilibrium between all pairs of outlier SNPs from EigenGWAS Eigenvector 1 and malaria BSLMM analysis. We thank the many people who helped us in the field, especially JC Illera.

Local governments in the Canary Islands and Madeira provided accommodation; the Spanish Environment Ministry, the Canary government and the Natural Park of Madeira gave permission to perform the sampling work, and the Portuguese Navy helped with transport. We thank the associate editor and two anonymous reviewers for their comments, which helped us to improve the manuscript. This work was supported by a Natural Environment Research Council (NERC) studentship to C.A., a Norwich Research Park Science Links grant to D.S.R., M.C., and L.G.S., a Leverhulme Fellowship to T.B., and a BBSRC fellowship and British Ecological Society Large Research Grant to L.G.S. RAD library preparation was supported by NERC at the NERC Biomolecular Analysis Facility in Sheffield (NBAF‐S), and sequenced by Edinburgh Genomics, The University of Edinburgh. Edinburgh Genomics is partly supported through core grants from NERC (R8/H10/56), the Medical Research Council (MR/K001744/1) and the Biotechnology and Biological Sciences Research Council (BBSRC) (BB/J004243/1).

Next‐generation sequencing and library construction of the genome was delivered via the BBSRC National Capability in Genomics (BB/J010375/1) at the Earlham Institute by members of the Platforms and Pipelines Group, strategically funded by BBSRC, Institute Strategic Programme Grant (BB/J004669/1).

How to download yvert et tellier files to my device?. 1. Click download file button or Copy yvert et tellier URL which shown in textarea when you clicked file title, and paste it into your browsers address bar. If file is multipart don't forget to check all parts before downloading!. 2. In next page click regular or free download and wait certain amount of time (usually around 30 seconds) until download button will appead.

3. Click it and That's it, you're done amigo! Yvert et tellier download will begin.