How Can Genomic Tools Contribute to the Conservation of...

72
International Journal of Genomics How Can Genomic Tools Contribute to the Conservation of Endangered Organisms Guest Editors: Cino Pertoldi, Ettore Randi, Aritz Ruiz‑González, Philippine Vergeer, and Joop Ouborg

Transcript of How Can Genomic Tools Contribute to the Conservation of...

  • International Journal of Genomics

    How Can Genomic Tools Contribute to the Conservation of Endangered Organisms

    Guest Editors: Cino Pertoldi, Ettore Randi, Aritz Ruiz‑González, Philippine Vergeer, and Joop Ouborg

  • How Can Genomic Tools Contribute tothe Conservation of Endangered Organisms

  • International Journal of Genomics

    How Can Genomic Tools Contribute tothe Conservation of Endangered Organisms

    Guest Editors: CinoPertoldi, Ettore Randi, Aritz Ruiz-González,Philippine Vergeer, and Joop Ouborg

  • Copyright © 2016 Hindawi Publishing Corporation. All rights reserved.

    This is a special issue published in “International Journal of Genomics.” All articles are open access articles distributed under the CreativeCommons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the originalwork is properly cited.

  • Editorial Board

    Jacques Camonis, FranceShen Liang Chen, TaiwanPrabhakara V. Choudary, USAMartine A. Collart, SwitzerlandSoraya E. Gutierrez, ChileM. Hadzopoulou-Cladaras, Greece

    Sylvia Hagemann, AustriaHenry Heng, USAEivind Hovig, NorwayGiuliana Napolitano, ItalyFerenc Olasz, HungaryElena Pasyukova, Russia

    Graziano Pesole, ItalyGiulia Piaggio, ItalyMohamed Salem, USABrian Wigdahl, USAJinfa Zhang, USA

  • Contents

    How Can Genomic Tools Contribute to the Conservation of Endangered OrganismsCino Pertoldi, Ettore Randi, Aritz Ruiz-González, Philippine Vergeer, and Joop OuborgVolume 2016, Article ID 4712487, 2 pages

    Novel Graphical Analyses of Runs of Homozygosity among Species and Livestock BreedsLaura Iacolina, Astrid V. Stronen, Cino Pertoldi, Małgorzata Tokarska, Louise S. Nørgaard, Joaquin Muñoz,Anders Kjærsgaard, Aritz Ruiz-Gonzalez, Stanisław Kami'nski, and Deirdre C. PurfieldVolume 2016, Article ID 2152847, 8 pages

    Using Genome-Wide SNP Discovery and Genotyping to Reveal the Main Source of PopulationDifferentiation in Nothofagus dombeyi (Mirb.) Oerst. in ChileRodrigo Hasbún, Jorge González, Carolina Iturra, Glenda Fuentes, Diego Alarcón, and Eduardo RuizVolume 2016, Article ID 3654093, 10 pages

    Integrating Genomic Data Sets for Knowledge Discovery: An Informed Approach to Management ofCaptive Endangered SpeciesKristopher J. L. Irizarry, Doug Bryant, Jordan Kalish, Curtis Eng, Peggy L. Schmidt, Gini Barrett,and Margaret C. BarrVolume 2016, Article ID 2374610, 12 pages

    A Quantitative Genomic Approach for Analysis of Fitness and Stress Related Traits in a DrosophilamelanogasterModel PopulationPalle Duun Rohde, Kristian Krag, Volker Loeschcke, Johannes Overgaard, Peter Sørensen,and Torsten Nygaard KristensenVolume 2016, Article ID 2157494, 11 pages

    TheMicrobiome of Animals: Implications for Conservation BiologySimon Bahrndorff, Tibebu Alemu, Temesgen Alemneh, and Jeppe Lund NielsenVolume 2016, Article ID 5304028, 7 pages

    Differential Methylation of Genomic Regions Associated with Heteroblasty Detected by M&MAlgorithm in the Nonmodel Species Eucalyptus globulus Labill.Rodrigo Hasbún, Carolina Iturra, Soraya Bravo, Boris Rebolledo-Jaramillo, and Luis ValledorVolume 2016, Article ID 4395153, 7 pages

    TheUse of Genomics in ConservationManagement of the Endangered VisayanWarty Pig (Sus cebifrons)Rascha J. M. Nuijten, Mirte Bosse, Richard P. M. A. Crooijmans, Ole Madsen, Willem Schaftenaar,Oliver A. Ryder, Martien A. M. Groenen, and Hendrik-Jan MegensVolume 2016, Article ID 5613862, 9 pages

  • EditorialHow Can Genomic Tools Contribute to the Conservation ofEndangered Organisms

    Cino Pertoldi,1 Ettore Randi,2 Aritz Ruiz-González,3

    Philippine Vergeer,4 and Joop Ouborg5

    1Department of Chemistry and Bioscience, Aalborg University, Aalborg Zoo, Mølleparkvej 63, 9000 Aalborg, Denmark2Istituto Superiore per la Protezione e la Ricerca Ambientale ISPRA, Roma, Italy3Department of Zoology and Animal Cell Biology, University of the Basque Country, UPV/EHU, Vitoria-Gasteiz, Spain4Nature Conservation and Plant Ecology Group, Wageningen University, Wageningen, Netherlands5Experimentele Planten Ecologie, IWWR, Radboud Universiteit, Nijmegen, Netherlands

    Correspondence should be addressed to Cino Pertoldi; [email protected]

    Received 19 October 2016; Accepted 19 October 2016

    Copyright © 2016 Cino Pertoldi et al. This is an open access article distributed under the Creative Commons Attribution License,which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

    Conservation biologists have realized the urgent need forgenomic tools and interdisciplinary approaches to helpunderstand the loss of biodiversity. Conservation genetics hasbeen successful in highlighting the roles of evolutionary andpopulation genetics for biodiversity conservation, yet severalcritical issues remain. Rapid development of new sequencingtechniques means conservation genomics can now helpanswer some of the crucial issues that conservation geneticswas able to highlight but not to resolve.These include a betterunderstanding of outbreeding and inbreeding depression andthe extent to which estimates of heterozygosity accuratelyreflect quantitative trait variation and fitness. Conservationgenomics can help resolve taxonomic uncertainties, con-tribute to designation of evolutionary significant units andecotypes, and identify contemporary versus historic patternsof hybridization. The definition of a species is neverthelessa normative concept and cannot be resolved by genomicsalone. All living organisms are faced with amultitude of chal-lenges in their natural environment, such as climate change,diseases, predation, competition, and habitat disturbance. Inthe short term, animals and plants can acclimatize to shiftingenvironmental conditions by developing and expressing par-ticular traits in response to local environmental conditions(phenotypic plasticity). Organisms can also react to theshifting environment by dispersal; however this option isnot always available when, for example, the landscape is toofragmented. The last type of response is evolution via genetic

    selection leading to adaptation.The persistence of species andpopulations depends however on the initial response to theshifts in the environment.

    The genome of a species contains signatures of theseresponses that may be studied with genetic markers. We areinterested in filling gaps in the knowledge about past demo-graphic history of organisms and the factors that ultimatelyshape genomic variation in populations. For this we use thevery latest, innovative techniques in next generation sequenc-ing. The incorporation of technological developments inmolecular biology and the ongoing development of genomictools, like SNPs and next generation sequencing, andgenomic-based approaches, like full genome scans and gene-expression pattern analysis, make it possible to address ques-tions that until now were hard to tackle. There is an urgentneed for empirical studies on nonmodel organisms whichcan contribute to the emerging disciplines of populationgenomics and landscape genomics. Such studies are neces-sary if wewant to showhow these advances inmolecular tech-niques and approaches allow conservation genetics to makea big leap forward. Incorporating genetics and genomics intonature conservation will highlight the following:

    (i) Genomic consequences of inbreeding(ii) Inbreeding by environment interaction(iii) Genomic and epigenomic consequences of outbreed-

    ing

    Hindawi Publishing CorporationInternational Journal of GenomicsVolume 2016, Article ID 4712487, 2 pageshttp://dx.doi.org/10.1155/2016/4712487

    http://dx.doi.org/10.1155/2016/4712487

  • 2 International Journal of Genomics

    (iv) Genomic and epigenomic mechanisms of phenotypicplasticity

    (v) Transcriptome, metabolomics, and proteomic tech-niques applied to conservation biology

    (vi) The emerging discipline of landscape genomics;detection of signature of selection using genomictechniques

    All the results from the accepted papers have greatlycontributed to the special issue. To optimize the use oflimited resources in conservation biology and conservationgenomics and minimize the need to sample and disturb wildspecies (often those most in need of investigation are themost sensitive to disturbance), we recommend increasedpublic sharing of resources including genetic markers, data,and analytical tools and improved professional recognitionfor the publication of genomic resources.

    Acknowledgments

    Special thanks also extend to all the authors and referees fortheir efforts in the special issue.

    Cino PertoldiEttore Randi

    Aritz Ruiz-GonzálezPhilippine Vergeer

    Joop Ouborg

  • Research ArticleNovel Graphical Analyses of Runs of Homozygosity amongSpecies and Livestock Breeds

    Laura Iacolina,1 Astrid V. Stronen,1 Cino Pertoldi,1,2 MaBgorzata Tokarska,3

    Louise S. Nørgaard,1 Joaquin Muñoz,1 Anders Kjærsgaard,1 Aritz Ruiz-Gonzalez,4

    StanisBaw KamiNski,5 and Deirdre C. Purfield6

    1Department of Chemistry and Bioscience, Aalborg University, Section of Biology and Environmental Engineering,Fredrik Bajers Vej 7H, 9220 Aalborg, Denmark2Aalborg Zoo, Mølleparkvej 63, 9000 Aalborg, Denmark3Mammal Research Institute Polish Academy of Sciences, Ul. Waszkiewicza 1, 17-230 Białowieża, Poland4Department of Zoology and Animal Cell Biology, University of the Basque Country UPV/EHU,C/Paseo de la Universidad 7, 01006 Vitoria-Gasteiz, Spain5Department of Animal Genetics, University of Warmia and Mazury in Olsztyn, 10-718 Olsztyn, Poland6Animal & Biosciences Department, Animal & Grassland Research and Innovation Centre, Teagasc, Moorepark, Fermoy,County Cork, Ireland

    Correspondence should be addressed to Laura Iacolina; [email protected]

    Received 14 December 2015; Accepted 28 September 2016

    Academic Editor: Graziano Pesole

    Copyright © 2016 Laura Iacolina et al. This is an open access article distributed under the Creative Commons Attribution License,which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

    Runs of homozygosity (ROH), uninterrupted stretches of homozygous genotypes resulting from parents transmitting identicalhaplotypes to their offspring, have emerged as informative genome-wide estimates of autozygosity (inbreeding). We used genomicprofiles based on 698K single nucleotide polymorphisms (SNPs) from nine breeds of domestic cattle (Bos taurus) and the Europeanbison (Bison bonasus) to investigate how ROH distributions can be compared within and among species. We focused on twolength classes: 0.5–15Mb to investigate ancient events and >15Mb to address recent events (approximately three generations). Foreach length class, we chose a few chromosomes with a high number of ROH, calculated the percentage of times a SNP appearedin a ROH, and plotted the results. We selected areas with distinct patterns including regions where (1) all groups revealed anincrease or decrease of ROH, (2) bison differed from cattle, (3) one cattle breed or groups of breeds differed (e.g., dairy versusmeat cattle). Examination of these regions in the cattle genome showed genes potentially important for natural and human-induced selection, concerning, for example, meat and milk quality, metabolism, growth, and immune function. The comparativemethodology presented here permits visual identification of regions of interest for selection, breeding programs, and conservation.

    1. Introduction

    Mating among closely related individuals can affect the fitnessof the progeny by increasing the inbreeding coefficient (F) [1]and therefore the probability that alleles at a locus, sampledrandomly in a population, are identical by descent (IBD) [2].The reduction in fitness can be due to the accumulation ofrecessive lethal genetic disorders, reduction of fertility, andlower adaptive potential [1, 3, 4].

    In wild living and captive populations, there is an urgentneed to reduce inbreeding and augment genetic diversity,and this can be achieved by implementing carefully plannedmating strategies. One possibility consists in reducing thelevel of inbreeding per generation and the response toselection (optimal contribution selection) [5].The estimationof F requires completeness and accuracy of the availablepedigree records, which are not always available, because ofmissing information or registration errors. When genotypes

    Hindawi Publishing CorporationInternational Journal of GenomicsVolume 2016, Article ID 2152847, 8 pageshttp://dx.doi.org/10.1155/2016/2152847

    http://dx.doi.org/10.1155/2016/2152847

  • 2 International Journal of Genomics

    are available a probabilistic approach can be utilized for thereconstruction of the pedigree. However, such an approachdoes not take into account the stochastic nature of recombi-nation [6]. New approaches based on the runs of homozy-gosity (ROH), which are DNA segments that harbour unin-terrupted stretches of homozygous genotypes, have shown tobe reliable estimates of autozygosity at the genome-wide level[7–9].

    In addition, the frequency and extent of ROH can be usedto estimate the time when the inbreeding event took place.Considering that recombination events break long chromo-some segments, it is assumed that long autozygous segmentsin an individual derive from a common recent ancestor,whereas shorter autozygous segments are indicating a remotecommon ancestor [10–12]. We should therefore expect thatthe longer the homozygous segments, the more recent theinbreeding. However, long ROH may also be explained by arecent event under strong selective pressure. ROH can thusbe used to identify the genomic signatures of recent and/orancient selective pressure, as shown by [9]. Additionally, fixedROH in all the individuals in a population could indicate pastselective events. Clearly, the presence of long ROH at rela-tively high frequency in a population could also indicate thepresence of genetic substructure, with consanguineous mat-ing occurring onlywithin some subpopulations [13]. ROHarealso affected by demographic events [8] and further inves-tigation should examine issues such as skewed reproductivesuccess.

    The objective of this study was to describe and comparethe distribution of ROH of different length in nine Bostaurus cattle breeds under different management practicesand selection histories. The same comparison was madeat the interspecific level by comparing the distribution ofthe ROH between the abovementioned cattle breeds andthe Lowland line of the European bison (Bison bonasus)from the Białowieża National Park (Poland). The Low-land line is highly inbred due to only seven founders[14].

    While previous investigations were exclusively based onthe count and sum of the number of ROH above a givenlength [9], in this paper we analysed the frequency of SNPsfalling within a ROH above and below an a priori chosenlength (15Mb) and we visualized the different distributionsacross populations. In addition, this graphical visualizationallows the identification of similarities and dissimilaritiesin the regions that can be used to investigate possibleadaptive/selective patterns.

    2. Material and Methods

    2.1. Genotypes and Quality Control. Genotypes consisting of777,972 single nucleotide polymorphisms (SNPs) from theBovineHD BeadChip (Illumina Inc., San Diego, CA) weregenerated for 891 sires of multiple breeds. Breeds representedinclude Angus (𝑛 = 39), Belgian Blue (𝑛 = 38), Charolais(𝑛 = 117), Friesian (𝑛 = 98), Hereford (𝑛 = 40), Holstein(𝑛 = 262),Holstein-Friesian crosses (𝑛 = 111), Limousin (𝑛 =128), and Simmental (𝑛 = 58) (data from [9]). Angus, Belgian

    Blue, and Hereford are primarily meat breeds; Friesian,Holstein, and Holstein-Friesian crosses are primarily dairybreeds, while Limousin, Simmental, and Charolais are usedfor both milk and meat. Forty European Lowland bison(Bison bonasus) from Białowieża National Park (Poland)were used for comparison. GenomeStudio� (Illumina Inc.,San Diego, CA) and accompanying guidelines from Illumina(http://www.illumina.com/Documents/products/technotes/technote infinium genotyping data analysis.pdf) were usedfor quality control. Total individual call rate in the bisonwas 0.99. For cattle, only biallelic SNPs on the 29 autosomeswere retained after removing all monomorphic SNPsacross breeds, filtering for Hardy Weinberg Equilibrium(𝑝 < 0.0001) within each breed separately and for call rates>90%. Final analyses were performed on 867 cattle and 40bison with 698,384 SNPs.

    2.2. Runs of Homozygosity. Following the approach in [9],ROHwere estimated usingPLINKv1.07 [15] andwere definedwithin a sliding window of 50 SNPs, in one SNP interval,across the genome. Up to one possible heterozygous genotypewas permitted and no more than two SNPs with missinggenotypes were allowed per window (see [9]).

    ROH were divided in seven length categories (1–5Mb,5–10Mb, 10–15Mb, 15–20Mb, 20–25MB, 25–30Mb, and>30Mb). For each ROH length category we summed allROH per animal and averaged this per cattle breed andfor the bison. In order to investigate the potential of ourapproach, we then focused on two length classes: from500Kb till 15Mb to investigate ancient events and >15Mbto address recent events. To select target chromosomes fordetailed analyses, we created Manhattan plots with SAS9.4 (SAS Institute Inc., Toronto, Canada) for both lengthclasses and selected the chromosomes accordingly. For thechosen chromosomes, we calculated the percentage of timesa SNP appeared in a ROH and plotted these results withSAS.

    2.3. Analyses of Genomic Regions in the Runs of Homozy-gosity. As an example for the methodology applied in thisstudy, we selected regions of the different chromosomesthat showed one of the following patterns (see Figure 2):(a) a simultaneous increase (or decrease) in the numberof SNPs in a ROH across all populations, as this patterncould possibly involve genes fundamental for the two speciesanalysed; (b) few populations showing an opposite patterncompared to the others, as this could comprise genes specificfor those populations; (c) different patterns between dairyandmeat breeds, as this could possibly concern regions underhuman-induced directional selection; (d) different patternsbetween bison and domestic cattle breeds, as this patternmay be related to traits important for survival in the wild;(e) a single domestic breed differentiating from the others,as this could relate to specific characteristics of that breed;(f) a long region with a high percentage of ROH, as thiscould be associated with recent selective events; (g) a shortregion with opposite trend within a longer homogeneousregion, to investigate what could have caused such anabrupt change in variability levels. Each region was screened

    http://www.illumina.com/Documents/products/technotes/technote_infinium_genotyping_data_analysis.pdfhttp://www.illumina.com/Documents/products/technotes/technote_infinium_genotyping_data_analysis.pdf

  • International Journal of Genomics 3

    0

    200

    400

    600

    800

    1000

    1200

    1400

    1600

    1–5 5–10 10–15 15–20 20–25 25–30 >30(Mb)

    Mea

    n (M

    b)

    AngusBelgian BlueCharolaisFriesianHereford

    HolsteinHolstein-FriesianLimousinSimmentalBison

    Figure 1: The mean sum of runs of homozygosity (ROH) pergenotyped individual, measured in Megabases (Mb) within eachpopulation, for each considered ROH length category.

    using NCBI (https://www.ncbi.nlm.nih.gov/) resources forthe presence of annotated genes (release 104) and informa-tion on their biological function and possible evolutionaryimportance.

    3. Results

    3.1. Runs of Homozygosity. TheEuropean bison exhibited thehighest mean sum of ROH in the length categories 1–5Mb,5-10Mb, and 10–15Mb compared to all the domestic breeds.Angus and Hereford also showed considerably higher meansums than other breeds in the categories 1–5Mb and 5–10Mb(see Figure 1).

    In the Manhattan plot for the length class between500Kb and 15Mb, chromosomes 2 and 3 showed a group ofextremely variable SNPs, while chromosomes 7, 14, and 16 hadthe highest density and frequencies of SNPs falling in a ROH(see Figure S1a in Supplementary Material available online athttp://dx.doi.org/10.1155/2016/2152847). We thus focused onthese chromosomes for subsequent analyses. For the ROH>15Mb, the Manhattan plot showed a more homogeneousdistribution but we selected chromosomes 6, 9, and 20 forsubsequent analyses (Figure S2a). In the plots based on ROH< 15Mb, we observed large regions of the bison genomewhere almost 100% of SNPs fell within a ROH (Figure S1b–f). The frequency of SNPs falling in a ROH > 15Mb waslower for all populations, in accordance with the smallernumber of ROH in this length category (Figure S2b–d).Additionally, the frequency of a SNP falling within a ROH in

    the bison was not higher than that observed in the domesticbreedswith a single exception on chromosome 9 (Figure S2c).On chromosome 20 the highest percentage of SNPs fallingwithin a ROH was detected in dairy cattle breeds (FigureS2d). No clear pattern was observed on chromosome 6(Figure S2b).

    3.2. Analyses of Genomic Regions in the Runs of Homozygosity.The in-depth analysis of 17 regions, selected from sevenchromosomes (i.e., 2, 3, 7, 9, 14, 16, and 20) led to the iden-tification of more than 300 annotated genes whose functionsvary considerably (see Table S1). The most frequent func-tionally characterised genes were those related to metabolicpathways, but we also observed genes related to disease andimmune function, growth, and reproduction. As an example,we review here a few of our observations in the selectedregions.

    In summary, pattern (a) were mainly related to metabolicpathways, involving several CD-, ATP-, and SLAM-familygenes (see Table S1) and olfactory receptors. Metabolic path-ways were the main genes observed in pattern (b). Pattern(c) was inconclusive for ROH < 15Mb. In pattern (f) (also anexample of (c)) ROH > 15Mb included genes related to milkand meat quality, growth, and metabolic disorders related toenergy unbalanced consumption. Patterns (d) were locatedin portions of the chromosomes poorly described, with theonly exception being the long region on chromosome 9,where a high number of ROH > 15Mb was observed (FigureS2c). In addition to the metabolism and disease related geneswidely encountered in all the screened regions, we reportthe presence of genes related to olfactory perception, obesity,growth, and sperm malformation in this region. In pattern(e), we observed a region (Figure 2(e)) where the Simmentalshowed higher variability than the other breeds. Here, genesinvolved were related to fat thickness and colour, growth, andsperm functionality. In pattern (f), where Hereford showedextremely high frequency values of SNP falling within aROH and the Belgian Blue extreme variability (with theother breeds in between; Figure S1f, near 45000000), thegenes observed were mainly related to the codification ofproteins involved in sugar transport and assimilation atcellular level. In pattern (g) we observed genes involvedin cortisol pathways and sweet perception, regulation ofhost response to virus infection, and regulatory function inovulation.

    4. Discussion

    Our findings revealed several chromosomes with a highnumber of ROH, and most results concerned ROH <15Mb. Upon closer inspection of selected chromosomes, weobserved genes potentially important for natural and human-induced selection, concerning, for example, meat and milkquality, metabolism, growth, and immune function. Hence,the ROH approach appears informative for evaluating andcomparing species and population history and evaluatingpossible patterns of adaptation.

    We observed comparatively few results for ROH > 15Mb,the longer regions that are likely to reflect recent inbreeding

    https://www.ncbi.nlm.nih.gov/http://dx.doi.org/10.1155/2016/2152847)

  • 4 International Journal of Genomics

    % o

    f tim

    e SN

    P ap

    pear

    ed in

    a RO

    H

    010203040506070 8090

    100

    SNP position8 9 10 11

    BreedBisonHEAABBSI

    CHLMHOLFRHOFR

    ×106

    (a)

    % o

    f tim

    e SN

    P ap

    pear

    ed in

    a RO

    H

    010203040506070 8090

    100

    SNP position458 462460 464 468 472466 470 474

    ×105

    BreedBisonHEAABBSI

    CHLMHOLFRHOFR

    (b)

    % o

    f tim

    e SN

    P ap

    pear

    ed in

    a RO

    H

    010203040506070 8090

    100

    SNP position121 122 123 124

    ×106

    BreedBisonHEAABBSI

    CHLMHOLFRHOFR

    (c)

    % o

    f tim

    e SN

    P ap

    pear

    ed in

    a RO

    H

    010203040506070

    8090

    100

    SNP position1084 1088 1092 1096 1100 1104

    ×105

    BreedBisonHEAABBSI

    CHLMHOLFRHOFR

    (d)

    % o

    f tim

    e SN

    P ap

    pear

    ed in

    a RO

    H

    010203040506070 8090

    100

    SNP position240 242 244 246 248 250 252 254 256 258

    ×105

    BreedBisonHEAABBSI

    CHLMHOLFRHOFR

    (e)

    % o

    f tim

    e SN

    P ap

    pear

    ed in

    a RO

    H

    0123456789

    1011

    SNP position20 30 40 50

    ×106

    BreedBisonHEAABBSI

    CHLMHOLFRHOFR

    (f)

    Figure 2: Continued.

  • International Journal of Genomics 5

    % o

    f tim

    e SN

    P ap

    pear

    ed in

    a RO

    H

    010203040506070

    8090

    100

    SNP position508 512 516 520510 514 518 522

    ×105

    BreedBisonHEAABBSI

    CHLMHOLFRHOFR

    (g)

    Figure 2: Examples of the investigated patterns. (a) A simultaneous increase (or decrease) in the number of SNPs in a ROH across allpopulations, as this pattern could possibly involve genes fundamental for the two species analysed (chromosome 3); (b) few populationsshowing an opposite pattern compared to the others, as this could comprise genes specific for those populations (chromosome 7); (c) differentpatterns between dairy and meat breeds, as this could possibly concern regions under human-induced directional selection (chromosome2); (d) different patterns between bison and domestic cattle breeds, as this pattern may be related to traits important for survival in thewild (chromosome 3); (e) a single domestic breed differentiating from the others, as this could relate to specific characteristics of that breed(chromosome 14); (f) a long region with a high percentage of ROH, as this could be associated with recent selective events (chromosome 20);(g) a short region with opposite trend within a longer homogeneous region, to investigate what could have caused such an abrupt change invariability levels (chromosome 7).

    [9, 11]. Our results may thus suggest relatively limited recentinbreeding in the cattle breeds included in the study, althoughthe many shorter ROH could indicate a lower 𝑁E in thepast [16]. For the European bison, however, large regions ofthe genome had a 100% (or near 100%) frequency of SNPsfalling within a ROH.This suggests high levels of inbreeding,which is consistentwith earlier studies and knownpopulationhistory involving a severe bottleneck [17, 18]. However, evenlimited inbreeding can cause detrimental effects [1, 19] andshould be monitored. Earlier studies across species havesuggested that ROH > 16Mb may be considered as recentinbreeding [11, 16]. Analyses of cattle breeds report ROH> 16Mb as the expected mean after approximately threegenerations since the most recent common ancestor, whereasautozygosity due to more distant common ancestors will notbe captured by this measure [11]. For an in-depth assessmentof inbreeding, it may be necessary to investigate differentROH length classes considering the history of the organismsunder study. For example, comparisons between wild anddomestic species may show different patterns than native andcommercial livestock in terms of recent and/or past historiesof inbreeding. Consequently, ROH length classes should beassessed on a case by case basis with exploratory analysesinformed, where possible, by the history of the species understudy.

    Variation in sample size and𝑁E may have influenced theresults. Our comparison of, for example, Belgian Blue (𝑛 =38) and Holstein (𝑛 = 262) should therefore be interpretedwith caution. Other important factors that may play a roleare differences in breed genetic diversity. McTavish et al. [20]reported observed heterozygosity for several breeds includedin our study based on 50K SNP markers. Among the breedsthat showed distinct ROH patterns in our study, they notethat Simmental showed a heterozygosity of 0.28 (𝑛 = 77), theBelgian Blue 0.30 (𝑛 = 4), the Hereford 0.29 (𝑛 = 98), and theHolstein 0.30 (𝑛 = 85). Furthermore, the value for Limous-in was 0.29 (𝑛 = 100) and for Charolais was 0.31 (𝑛 = 53). Al-though these values are similar despite variable sample size,among- and within-breed variation in genetic diversity couldaffect ROH results and their interpretation andmay thereforecomplicate our comparison of cattle breeds and Europeanbison.

    Angus and Hereford breeds, together with bison, showhigh mean sum of ROH in the length class 1–10Mb, whichmay be a result of ancestral relatedness owing to smallfounder populations and isolated origins [11]. In particular,the ROH for the bison is extremely high for the intervals 1–5Mb and 5–10Mb with several regions that are completelyfixed.This appears consistent with an estimated𝑁E of 23 anda total of seven founders for the European bison’s Lowland

  • 6 International Journal of Genomics

    population [18]. In comparison, a recent survey presentedconsiderably larger but variable census population size (𝑁C)and 𝑁E for some of the cattle breeds included in our study[21]. For Aberdeen Angus, they reported 𝑁C > 10M and𝑁E of 136. For Holstein, 𝑁C was >65M and 𝑁E was 99,whereas for Limousin, 𝑁C was >4M and 𝑁E was 174. Theremay thus be considerable differences in population historyamong breeds and also for breeds within the same group(such as meat production), which could have affected ourresults.

    We observed genes grouped into various functionalcategories. The types of genes observed may reveal adaptivepatterns and indicate human-induced and/or natural selec-tion, for example, in cases of genes linked to growth andimmunity where the first is likely to be human-modifiedand the second is subject to stronger natural selection. Ourresults also highlight the need to consider potential conflictsbetween these two sources of selection. For example, wenoted a gene implicated in ketosis (region F, chromosome20), a metabolic disorder that occurs in cattle when energydemands such as high milk production exceed energy intakeand result in a negative energy balance. Strong directionalselection for high-performance characteristics such as highmilk yield may therefore have implications for animal healthand welfare, life expectancy, and the ethical dimensions ofanimal breeding to cope with their living environments (see,e.g., [22, 23]).

    4.1. Applications. The ROH approach seems informativefor investigating selection and evolutionary histories acrossa range of different populations, including wild/domesticspecies, native/commercial livestock, and commercial breedsof various kinds (e.g., cattle breeds for milk or meat, sheepbreeds for meat or wool). Our study compared cattle withone related wild species, the European bison. However, thisspecies is highly inbred and has low genetic diversity [18].Study of other wild-domestic species pairs may thereforeprovide a more nuanced picture of genomic regions underselection, for example, in domestic pigs and wild boar, orcaptive and free-living populations of the wild boar (e.g.,[24]), thus taking advantage of recent developments in high-density genomic arrays to investigate domestic and wildspecies (e.g., [25]).

    The results of our analyses may also suggest applicationsfor genetic rescue. This could include key genetic regions ofhigh variability observed in one breed, which could be trans-ferred to one or more other populations, for example, relatedto immune system function or tolerance to environmentalfactors such as heat, parasites, and infectious disease [26,27]. Moreover, genes related to growth may have importantapplications for animal breeding and could be introduced tonew breeds to enhance both genetic variation and production[28]. Further research may also help clarify the extent towhich selection for rapid growthmight conflict with selectionfor meat quality, which may be relevant to conservationmanagement and breeding for both commercial and nativelivestock breeds (e.g., [29]).

    It will be important to establish whether ROH are underselection. If a ROH is not under selection, its length should

    normally decrease with every generation as the expectedlength of autozygous segments identical by descent fol-lows an exponential distribution with mean equal to 0.5𝑔Morgans, where 𝑔 is the number of generations since thecommon ancestor [30]. Conversely, a ROH could containrecessive variants that are expressed in the autozygous state.These variants are known to cause various genetic diseasesin humans as a result of specific mutations (e.g., phenylke-tonuria, Tay-Sachs disease, and cystic fibrosis) and may alsobe involved in complex diseases such as heart and liverdiseases and diabetes [31].

    For livestock, the incidence of disease associated withintensive production has increased among several breeds[32], such as Holstein and Jersey [33–35]. Additionally,important traits, such as adaptation to low-quality foodresources, parasites, and tolerance to disease and temperaturefluctuations may be found mostly in native breeds [36]. Animportant aspect of theROHassessmentwill be identificationof genetic variants with applications for genetic rescue, whichcould benefit both native and commercial breeds [28] toincrease robustness and tolerance to environmental variation[27, 36].

    4.2. Possible Limiting Factors. Ascertainment bias could haveaffected the comparison of ROH between different species(here cattle and bison) [37]. Moreover, our observations arenecessarily incomplete, as there are still large regions of thegenome that have not been fully described, as testified by thehigh number of uncharacterised genes we encountered in ourscreening (see Table S1). However, key genomic regions canbe noted for further research, which also helps identify high-priority areas of the genome for future study.

    5. Conclusions

    The comparative methodology presented here permits visualidentification of regions of interest, which could be of valuefor selection and breeding programs. The ROH approachoffers several immediate applications. Firstly, breeding strate-gies may be improved by reduction in ROH that are actingto reduce genomic diversity. Such a strategy could be usefulwhere genomic regions have lost important diversity or beenaccidentally fixed, for example, as a consequence of a pop-ulation bottleneck and/or founder effect. Further, the ROHapproach has implications for genetic rescue and the designof breeding strategies for populations at risk. The presenceof ROH at intermediate frequency in a population mayindicate heterogeneity of the𝑁E in different genomic regions.Accordingly, a breeding strategy based onmaximising𝑁E fora population could produce an increase of𝑁E for some chro-mosomal regions and a reduction in others. This situationcould complicate the design of a long-term protocol becauseof the risk of fixation of certain genes and loss of geneticdiversity. Human-driven breeding could also overwhelmnatural selective pressures, especially for populations mainlygoverned by genetic drift due to the small 𝑁E. It is thereforenecessary to balance various considerations for long-termconservation breeding, and information from ROH can help

  • International Journal of Genomics 7

    pinpoint important genomic regions even if we do not, at themoment, have a complete understanding of their function.

    Competing Interests

    The authors declare that they have no competing interests.

    Authors’ Contributions

    Laura Iacolina and Astrid V. Stronen contributed equally.

    Acknowledgments

    Cino Pertoldi was supported by a grant from Danish NaturalScience Research Council (Grant nos. 11-103926, 09-065999,and 95095995), the Carlsberg Foundation (Grant no. 2011-01-0059), and the Aalborg Zoo Conservation Foundation(AZCF). Laura Iacolina has received funding from the Euro-pean Union’s Horizon 2020 research and innovation pro-gramme under the Marie Sklodowska-Curie Action (GrantAgreement no. 656697). Astrid V. Stronen received fundingfrom the Danish Natural Science Research Council (Postdoc-toral Grant 1337-00007).

    References

    [1] D. W. Bjelland, K. A. Weigel, N. Vukasinovic, and J. D.Nkrumah, “Evaluation of inbreeding depression in Holsteincattle using whole-genome SNP markers and alternative mea-sures of genomic inbreeding,” Journal of Dairy Science, vol. 96,no. 7, pp. 4697–4706, 2013.

    [2] S. Wright, “Coefficients of inbreeding and relationship,” TheAmerican Naturalist, vol. 56, no. 645, pp. 330–338, 1922.

    [3] O. González-Recio, E. López de Maturana, and J. P. Gutiérrez,“Inbreeding depression on female fertility and calving ease inspanish dairy cattle,” Journal of Dairy Science, vol. 90, no. 12, pp.5744–5752, 2007.

    [4] S. Sommer, “The importance of immune gene variability(MHC) in evolutionary ecology and conservation,” Frontiers inZoology, vol. 2, article 16, 2005.

    [5] G. Gandini, A. Stella,M.Del Corvo, andG. B. Jansen, “Selectionwith inbreeding control in simulated young bull schemes forlocal dairy cattle breeds,” Journal of Dairy Science, vol. 97, no.3, pp. 1790–1798, 2014.

    [6] R. McQuillan, A.-L. Leutenegger, R. Abdel-Rahman et al.,“Runs of homozygosity in European populations,” The Ameri-can Journal of Human Genetics, vol. 83, no. 3, pp. 359–372, 2008.

    [7] J. Gibson, N. E. Morton, and A. Collins, “Extended tracts ofhomozygosity in outbred human populations,” Human Molec-ular Genetics, vol. 15, no. 5, pp. 789–795, 2006.

    [8] M. Bosse, H.-J. Megens, O. Madsen et al., “Regions of homozy-gosity in the porcine genome: consequence of demography andthe recombination landscape,” PLoS Genetics, vol. 8, no. 11,Article ID e1003100, 2012.

    [9] D. C. Purfield, D. P. Berry, S. McParland, and D. G. Bradley,“Runs of homozygosity and population history in cattle,” BMCGenetics, vol. 13, article 70, 2012.

    [10] M. Ferencakovic, E. Hamzic, B. Gredler, I. Curik, and J. Sölkner,“Runs of homozygosity reveal genomewide autozygosity in the

    Austrian Fleckvieh cattle,” Agriculturae Conspectus Scientificus,vol. 76, no. 4, pp. 325–328, 2011.

    [11] M. Ferenčaković, E. Hamzić, B. Gredler et al., “Estimates ofautozygosity derived from runs of homozygosity: empiricalevidence from selected cattle populations,” Journal of AnimalBreeding and Genetics, vol. 130, no. 4, pp. 286–293, 2013.

    [12] M. Ferenčaković, J. Sölkner, and I. Curik, “Estimating autozy-gosity from high-throughput information: effects of SNP den-sity and genotyping errors,”Genetics Selection Evolution, vol. 45,no. 1, article 42, 2013.

    [13] L. Iacolina, M. Scandura, D. J. Goedbloed et al., “Genomicdiversity and differentiation of a managed island wild boarpopulation,” Heredity, vol. 116, no. 1, pp. 60–67, 2016.

    [14] H. M. Slatis, “An analysis of inbreeding in the European bison,”Genetics, vol. 45, pp. 275–287, 1960.

    [15] S. Purcell, B. Neale, K. Todd-Brown et al., “Plink: a tool setfor whole-genome association and population-based linkageanalyses,” American Journal of Human Genetics, vol. 81, no. 3,pp. 559–575, 2007.

    [16] M. Kirin, R. McQuillan, C. S. Franklin, H. Campbell, P. M.McKeigue, and J. F. Wilson, “Genomic runs of homozygosityrecord population history and consanguinity,” PLoS ONE, vol.5, no. 11, Article ID e13996, 2010.

    [17] K. Oleński, M. Tokarska, D. M. Hering et al., “Genome-wideassociation study for posthitis in the free-living population ofEuropean bison (Bison bonasus),” Biology Direct, vol. 10, no. 2,2015.

    [18] M. Tokarska, C. Pertoldi, R. Kowalczyk, and K. Perzanowski,“Genetic status of the European bison Bison bonasus afterextinction in the wild and subsequent recovery,” MammalReview, vol. 41, no. 2, pp. 151–162, 2011.

    [19] J. E. Pryce, M. Haile-Mariam, M. E. Goddard, and B. J. Hayes,“Identification of genomic regions associated with inbreedingdepression in Holstein and Jersey dairy cattle,” Genetics Selec-tion Evolution, vol. 46, article 71, 2014.

    [20] E. J. McTavish, J. E. Decker, R. D. Schnabel, J. F. Taylor, andD. M. Hillis, “New World cattle show ancestry from multipleindependent domestication events,” Proceedings of the NationalAcademy of Sciences of the United States of America, vol. 110, no.15, pp. E1398–E1406, 2013.

    [21] The Bovine HapMap Consortium, “Genome-wide survey ofSNP variation uncovers the genetic structure of cattle breeds,”Science, vol. 324, no. 5926, pp. 528–532, 2009.

    [22] W. M. Muir, H.-W. Cheng, and C. Croney, “Methods to addresspoultry robustness and welfare issues through breeding andassociated ethical considerations,” Frontiers in Genetics, vol. 5,article 407, 2014.

    [23] W. M. Rauw and L. Gomez-Raya, “Genotype by environmentinteraction and breeding for robustness in livestock,” Frontiersin Genetics, vol. 6, article 310, 2015.

    [24] A. Canu, S. Costa, L. Iacolina, P. Piatti, M. Apollonio, and M.Scandura, “Are captive wild boar more introgressed than free-ranging wild boar? Two case studies in Italy,” European Journalof Wildlife Research, vol. 60, no. 3, pp. 459–467, 2014.

    [25] V. R. Kharzinova, A. A. Sermyagin, E. A. Gladyr, I. M. Okhlop-kov, G. Brem, and N. A. Zinovieva, “A study of applicability ofSNP chips developed for bovine and ovine species to whole-genome analysis of reindeer Rangifer tarandus,” Journal ofHeredity, vol. 106, no. 6, pp. 758–761, 2015.

  • 8 International Journal of Genomics

    [26] F. Alpay, Y. Zare, M. H. Kamalludin et al., “Genome-wideassociation study of susceptibility to infection by Mycobac-terium avium subspecies paratuberculosis in Holstein cattle,”PLoS ONE, vol. 9, no. 12, Article ID e111704, 2014.

    [27] L. R. Porto-Neto, A. Reverter, K. C. Prayaga et al., “The geneticarchitecture of climatic adaptation of tropical cattle,” PLoSONE,vol. 9, no. 11, Article ID e113284, 2014.

    [28] T.N.Kristensen,A.A.Hoffmann,C. Pertoldi, andA.V. Stronen,“What can livestock breeders learn from conservation geneticsand vice versa?” Frontiers in Genetics, vol. 5, article 38, 2015.

    [29] G. Holló, K. Nuernberg, T. Somogyi, I. Anton, and I. Holló,“Comparison of fattening performance and slaughter value oflocalHungarian cattle breeds to international breeds,”Archiv fürTierzucht, vol. 55, no. 1, pp. 1–12, 2012.

    [30] D. P. Howrigan, M. A. Simonson, and M. C. Keller, “Detectingautozygosity through runs of homozygosity: a comparison ofthree autozygosity detection algorithms,” BMC Genomics, vol.12, article 460, 2011.

    [31] L. B. Andrews, J. E. Fullarton, N. A. Holtzman, and A. G.Motulsky, Assessing Genetic Risks: Implications for Health andSocial Policy, National Academy Press, Washington, DC, USA,1994.

    [32] I. Medugorac, C. E. Veit-Kensch, J. Ramljak et al., “Conserva-tion priorities of genetic diversity in domesticated metapopu-lations: a study in taurine cattle breeds,” Ecology and Evolution,vol. 1, no. 3, pp. 408–420, 2011.

    [33] J. S. Agerholm, C. Bendixen, O. Andersen, and J. Arnbjerg,“Complex vertebral malformation in Holstein calves,” Journalof Veterinary Diagnostic Investigation, vol. 13, no. 4, pp. 283–289,2001.

    [34] P. M. VanRaden, K. M. Olson, D. J. Null, and J. L. Hutchison,“Harmful recessive effects on fertility detected by absence ofhomozygous haplotypes,” Journal of Dairy Science, vol. 94, no.12, pp. 6153–6161, 2011.

    [35] C. Charlier, J. S. Agerholm, W. Coppieters et al., “A deletion inthe bovine FANCI gene compromises fertility by causing fetaldeath and brachyspina,” PLoS ONE, vol. 7, no. 8, Article IDe43085, 2012.

    [36] I. Hoffmann, “Climate change and the characterization, breed-ing and conservation of animal genetic resources,” AnimalGenetics, vol. 41, no. 1, pp. 32–46, 2010.

    [37] L. K. Matukumalli, C. T. Lawley, R. D. Schnabel et al., “Devel-opment and characterization of a high density SNP genotypingassay for cattle,” PLoS ONE, vol. 4, no. 4, Article ID e5350, 2009.

  • Research ArticleUsing Genome-Wide SNP Discovery and Genotyping toReveal the Main Source of Population Differentiation inNothofagus dombeyi (Mirb.) Oerst. in Chile

    Rodrigo Hasbún,1 Jorge González,1 Carolina Iturra,1 Glenda Fuentes,2

    Diego Alarcón,2,3 and Eduardo Ruiz2

    1Departamento de Silvicultura, Facultad de Ciencias Forestales, Universidad de Concepción, 4070386 Concepción, Chile2Departamento de Botánica, Facultad de Ciencias Naturales y Oceanográficas, Universidad de Concepción,4070386 Concepción, Chile3Instituto de Ecologı́a y Biodiversidad (IEB), Facultad de Ciencias, Universidad de Chile, 7800003 Santiago, Chile

    Correspondence should be addressed to Rodrigo Hasbún; [email protected]

    Received 29 January 2016; Accepted 23 May 2016

    Academic Editor: Ettore Randi

    Copyright © 2016 Rodrigo Hasbún et al. This is an open access article distributed under the Creative Commons AttributionLicense, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properlycited.

    Within a woody plant species, environmental heterogeneity has the potential to influence the distribution of genetic variationamong populations through several evolutionary processes. In some species, a relationship between environmental characteristicsand the distribution of genotypes can be detected, showing the importance of natural selection as themain source of differentiation.Nothofagus dombeyi (Mirb.) Oerst. (Nothofagaceae) is an endemic tree species occurring both in Chile and in Argentina temperateforests. Postglacial history has been studied with chloroplast DNA and evolutionary forces shaping genetic variation patternshave been analysed with isozymes but fine-scale genetic diversity studies are needed. The study of demographic and selectionhistories in Nothofagus dombeyi requires more informative markers such as single nucleotide polymorphisms (SNP). Genotyping-by-Sequencing tools now allow studying thousands of SNP markers at reasonable prices in nonmodel species. We investigatedmore than 10K SNP loci for signatures of local adaptation and showed that interrogation of genomic resources can identify shiftsin genetic diversity and putative adaptive signals in this nonmodel woody species.

    1. Introduction

    In population genetics and conservation the big questionis genetic drift instead of natural selection. Definitely bothprocesses determine evolution, but genetic drift operates ran-domly and depends on effective population size while naturalselection proceeds nonrandomly and relies on environmentalvariables.The evolution towards hereditary adaptations to thecurrent environment is determined by natural selection andhas a direction; genetic drift instead is governed solely bychance. Consequently, drift acts on alleles, which generallyhave no phenotypic effect; instead selection favours certainalleles that increase fitness, reduce the unfavourable allelesfrequencies, and ignore neutral alleles [1]. Knowing what themain microevolutionary force is is very relevant for rational

    genetic management of threatened species, especially forspecies with geographical distribution severely fragmented[2].

    Approaches to addressing adaptive variation have beenincorporated into the definition of evolutionary significantunits [3, 4]. New technologies like Next Generation Sequenc-ing (NGS) and fine-scale GIS, coupled with advances incomputer hardware and software in the field of genomics[5–7], have allowed the development of new methods forcomprehensive evaluation of adaptive diversity [8, 9]. Thediscovery and genotyping of massive genetic markers arenow enabled by modern genomic tools at very low cost.This makes the study of adaptive genetic loci possible on awide range of species, which can facilitate the identificationof key biodiversity areas. Kirk and Freeland [10] reviewed

    Hindawi Publishing CorporationInternational Journal of GenomicsVolume 2016, Article ID 3654093, 10 pageshttp://dx.doi.org/10.1155/2016/3654093

    http://dx.doi.org/10.1155/2016/3654093

  • 2 International Journal of Genomics

    some of the applications of neutral versus adaptive markersin molecular ecology, discussed some of the advantagesthat can be obtained by supplementing studies of molecularecology with data from nonneutral molecular markers, andsummarized new methods that allow generating data fromloci under selection. Population genomic analyses requiremultilocus datasets from multiple populations and identifi-cation of nonneutral or adaptive loci by contrasting patternsof population divergence among genetic regions.

    Studies in nonmodel organisms have shown relativelybroad candidate genomic regions that are under selection, butit remains difficult to identify the genes (or the mutations)that are affected by selection. Increasing the density ofmarkers in genome scans is paramount to overcome thisproblem, and validating signals of selection from particulargenes using multiple methods should also help [4]. One ofthemost exciting developments in population genomics is thedevelopment of various reduced-representation protocols,collectively referred to as Genotyping-by-Sequencing (GBS),which allow sequencing of a subset of the genome throughselective amplification of restriction fragments [10].

    Nothofagus dombeyi (Mirb.) Oerst. (Nothofagaceae) isan endemic tree species occurring both in Chile and inArgentina temperate forests with a remarkably broad alti-tudinal and latitudinal distribution, across many differentecological gradients in the former [11]. The evergreen tree N.dombeyi is a pioneer species and constitutes an importantelement in the dynamics of South American forest. Itspostglacial history has been studied with chloroplast DNAand evolutionary forces shaping genetic variation patternshave been analysed with isozymes [12, 13]. However, genome-wide scan methods using thousands of markers to study arepresentative portion of the genome are needed.

    In this work, we assess GBS in the nonmodel woodyspecies N. dombeyi to develop high quantity of informativemarkers such as single nucleotide polymorphisms (SNP).Our aim is to determine the contribution of selection andmolecular adaptation to shaping genome-wide variation. Weexpect higher genetic population differentiation (for ecolog-ical separated localities) for adaptive SNP than neutral SNPif natural selection is the principal source of differentiation.Alternatively, if other sources of differentiation (mutation,genetic drift, and migration) are relevant, they will equallyaffect both types of SNP. Knowing the contribution ofselection effect to shaping genome variation patternswill havemany applications for biodiversity conservation, especiallyin endangered species, because neutral and adaptive geneticdiversity will likely have different impacts on long-termsurvival. In fact, in most cases, only adaptive diversity willallow a population to adapt to changing environmentalconditions.

    2. Materials and Methods

    2.1. Sampling Design

    2.1.1. Niche Modelling. To consider the remarkably broadaltitudinal and latitudinal distribution of this species, across

    many different ecological gradients, we used the methodproposed by Alarcón and Cavieres [14] to niche modellingof Nothofagus dombeyi in Chile using BIOMOD [15]. Eightvariables were selected from the WorldClim global climatedatabase [16] corresponding to the present climate condi-tions with a 30-arc-second grid resolution, with the leastcorrelation among them for the studied species range area.Four variables were related to energy constraints: (a) BIO2:mean diurnal temperature range; (b) BIO4: temperatureseasonality; (c) BIO5: maximum temperature in the warmestmonth; and (d) BIO6: minimum temperature in the coldestmonth. Other four variables were related to water availability:(e) BIO12: annual precipitation; (f) BIO15: precipitationseasonality; (g) BIO18: precipitation in the warmest quarter;and (h) BIO19: precipitation in the coldest quarter. Then,we projected the current and future distribution (year 2050)considering a conservative future climate projection CSIROB2A 2050 by using the tools of BIOMOD software. Further,we identified geographical areas with potential habitat loss,which should be identified as high priority in genetic conser-vation programs.

    2.1.2. Ecological Regions or Strata. Relatively homogeneousunits in ecological terms (strata) were defined from naturalpopulations of the species associated with the geographi-cal areas projected in its ecological niche modelling. TheCalinski-Harabasz criterion, which is a pseudo-𝐹 statistic asin ANOVA, was used to assess the best number of strataidentified by 𝐾-means partitioning [17].

    2.2. DNA Sample and Library Preparation

    2.2.1. Plant Material and Genomic DNA Isolation. Adult treeswith a diameter higher than 50 cm were sampled during thegrowing season of 2013-2014 from twenty-one sites coveringalmost the entire range of the species N. dombeyi in Chile.One to four sites were assigned to each stratum according toits superficies and 2 to 9 samples per site were taken (Table 1;Supplementary Figure 1 in Supplementary Material availableonline at http://dx.doi.org/10.1155/2016/3654093).

    DNA extraction was performed using a Qiagen DNeasyPlant kit (Qiagen Inc., USA). Lyophilized leaf tissue (20mg)was ground in a Precellys�24 (Precellys, USA) homogenizerwith two 1/4 ceramic spheres (MP BIOMEDICALS, USA)and AP1 buffer. The objective is grinding tissues and lysingcells prior to DNA extraction. DNA extraction protocol wasdone following themanufacturer instructions but elutionwasdone with 30 𝜇L (instead of 100 𝜇L) to increase the final DNAconcentration in the eluate (>100 ng/𝜇L). The integrity ofgenomic DNA was evaluated by agarose gel and quantifiedusing a Qubit fluorometer (Invitrogen, USA).

    2.2.2. Library Preparation and High-Throughput Sequencing.Library preparation and high-throughput sequencing wereperformed at University of Wisconsin Biotechnology Cen-ter (DNA Sequencing Facility). The GBS genomic librarypreparation was done following the protocol detailed byElshire et al. [18] with the methylation-sensitive restriction

  • International Journal of Genomics 3

    Table 1: Strata code and location, geographical coordinates, and sample size of the sampled individuals of Nothofagus dombeyi in Chile.

    Stratum Location (code) Latitude Longitude 𝑁1 Altos Lircay (AL) −35.599162 −71.044414 41 Antuco (AN) −37.343457 −71.615626 52 Ralco (RA) −37.925041 −71.575168 62 Termas de Tolhuaca (TT) −38.235047 −71.727552 93 Lago La Paloma (LP) −45.876213 −72.070813 43 El Machi (EM) −45.009553 −71.906872 53 Near Villa Amengual (NVAM) −45.008498 −71.908366 23 Villa Amengual (VAM) −45.007328 −71.911175 54 Nonguén (NO) −36.879745 −72.987923 54 Villa Las Araucarias (VA) −36.879774 −72.987981 64 Caramávida (CA) −36.879774 −72.987981 45 Mariquina (MA) −39.471811 −73.055799 45 Lago Riñihue (LRI) −39.480193 −73.048809 35 Fundo Llancahue (FL) −39.858686 −73.141572 35 Lago Neltume (LN) −39.859282 −73.141650 35 Parque Nacional Villarrica (PNV) −39.341073 −71.972351 55 Loncoche Interior (LI) −39.341820 −71.972232 35 Melipeuco (ME) −38.912949 −71.704088 36 Las Trancas (LT) −40.221101 −73.362430 56 Camino Osorno a Maicolpué (COM) −40.598321 −73.497015 47 Parque Nacional Puyehue (PNP) −40.737218 −72.306050 37 Lago Rupanco (LR) −40.736785 −72.302922 5

    enzyme ApeKI and 96 custom barcodes. Illumina high-throughput sequencing was conducted on an Illumina HiSeq2000 (Illumina, USA) using 100 bp single-end sequencingruns. The samples were sequenced across one Illumina lane.Base calling was performed in Casava v1.8.2 (Illumina, USA).

    2.3. Nonreference SNP Calling

    2.3.1. De Novo Identification of Loci/Alleles. Sequence resultswere analysed and SNP genotypes were assigned using theUNEAK (Universal Network Enabled Analysis Kit) GBSpipeline [19], which is part of the TASSEL 3.0 [20] bioin-formatic analysis package. This pipeline does not dependon a reference sequence, which is the actual case for N.dombeyi. SNP discovery is performed directly within pairsof matched sequence tags (unique sequence representing agroup of reads) and filtered through network analysis. Thenetwork filter trimmed reads to 64 bp to reduce the effectsof error sequencing and enabled efficient storage of data inbit format. SNP were assigned with default settings. Briefly,tags differing by a single nucleotide were retained as SNPand those with a minor allele frequency 0.05 were removedto minimize the impact of sequencing errors [19]. We used aminimum call rate of 0 and additional filters were applied innext steps.

    2.3.2. Post-SNP Calling Filters and Imputation. Given that weare trying to find SNP for population genetic analysis, we

    applied some filters to remove loci and individuals that con-tain very low levels of information prior to further analysis.We applied two functions of TASSEL 5.2 that removed allSNP (rows) and then samples (columns) containing 90% ormore “𝑁” values (indicating that neither allele is designated).These 𝑁s represent individuals where the allele cannot becalled from the sequence reads.This is because either no readis available at this site (for this individual) or the sequencequality is too low to be called.

    In order to cope with missing data, genotype imputationwas used to fill in the missing data and improve the powerof downstream analyses. We used LinkImpute implementedin TASSEL 5.2, a software package based on a 𝑘-nearestneighbour genotype imputation method, LD-𝑘NNi [21].Thisimputation method was designed specifically for nonmodelorganisms in which genomic resources are poorly developedand marker order is unreliable or unknown.

    2.4. Detection of Selection Footprints. To identify adaptiveSNP (putative loci under selection), we used LOSITAN [22].LOSITAN is a selection detection workbench based on theFst-outlier methods. We used 50,000 simulations, 0.99 forconfidence interval, false discovery rate of 0.05, mutationmodel “Infinite Alleles,” and the options “Neutral mean Fst”and “Force mean Fst”, which iteratively identify and removeFst outliers when calculating the global distribution of Fst.Our interest is in patterns of adaptation driven by environ-mental gradients; therefore we focused on outlier patterns

  • 4 International Journal of Genomics

    indicating divergent selection (Fst significantly higher thanneutral expectations).

    LOSITAN analyses were complemented with BayeScan2.1 [23] for estimating the posterior probability that a givenlocus is affected by selection. Briefly, prior odds of 100 wereused for identifying the top candidates of the selected loci anda total of 50,000 reversible-jumpMarkov Chain Monte Carlochains were run with a thinning interval of 10, following 20pilot runs of 5,000 iterations each, and a burn-in length of50,000. Loci were considered outliers with an FDR of 0.05.

    To confirm the adaptive SNP detected by previous meth-ods, the spatial analysis method (SAM) implemented inthe program Sam𝛽ada v0.5.1 [24] was used. We conductedthe analysis using the 10,109 SNP detected for N. dombeyiin Chile. Sam𝛽ada uses logistic regressions to model theprobability of presence of an allelic variant in a polymorphicmarker given the environmental conditions of the samplinglocations. Eight environmental variables previously describedin Section 2.1.1 were used (temperature related → BIO2,BIO4, BIO5, and BIO6; precipitation related → BIO12,BIO15, BIO18, and BIO19). Regarding genotypes, each ofthe states of a given SNP is considered independently asbinary presence/absence in each sample. Our biallelic SNPwere recoded as three distinct genotypes (AA, AB, and BB).A maximum likelihood approach is used to fit the modelsusing univariate analyses. Each model for a given genotypeis compared to a constant model, where the probability ofpresence of the genotype is the same at each location. Thestatistical significance threshold was set to 1% before applyingBonferroni correction. Significance was assessed with loglikelihood ratio (𝐺) tests [25] selecting loci/allele that testedhigher than the 99th percentile of the 𝐺 score distribution.

    2.5. Estimation of Genome-Wide Genetic Variation and Dif-ferentiation. Wemade all estimations in parallel with neutralSNP (10,109) and adaptive SNP.

    2.5.1. Basic Statistics of Genetic Variation. All the results wereobtained using the adegenet [26, 27] and hierFstat [28] Rpackages. Basic statistics were estimated including observedheterozygosity (Ho) and genetic diversity (Hs) within pop-ulation. Also, overall gene diversity (Ht) and corrected Ht(Htp), gene diversity among samples (Dst), and correctedDst (Dstp) were estimated. Fst and corrected Fst (Fstp) wereassessed as well as Fis followingNei [29] per overall loci. Dest,a measure of population differentiation as defined by Jost[30], was also calculated.The degree of genetic differentiationamong populations is expected to be low for neutral SNP buthighly divergent in SNP subject to directional selection.

    2.5.2. Population Structure. To describe the genetic biodiver-sity of a species, more important than diversity among indi-viduals is the diversity between groups of individuals. Firstwe analysed individual data to identify populations, or morelarge genetic clusters, and then we described these clustersby adegenet R package. To get a simplified picture of thegenetic diversity observed among individuals or populationswe used Principal Component Analysis (PCA). Discriminant

    Analysis of Principal Components (DAPC) function wasused to describe the relationships between these clusters.Themain results of DAPC were DAPC scatterplots.

    For each pair of strata, we computed pairwise Fst valueswith hierFstat R package. Principal Coordinates Analysis(PCoA) on Fst values was performed to detect major geneticclusters (𝐾) at individual level.

    2.5.3. Detecting Locus Contributions. In DAPC, the variablesactually analysed are principal components of a PCA. Load-ings of these variables are generally uninformative, since PCsthemselves do not all have straightforward interpretations.However, we can also compute contributions of the alleles,which can turn out to be very informative. In general, thereare many alleles and their contribution is best plotted fora single discriminant function at a time using adegenet Rpackage.

    3. Results and Discussion

    3.1. Niche Modelling and Strata Definition. Modelled areaof N. dombeyi presence under present climate conditionsis 88,174 km2 and in the future (year 2050) under dispersalconstraints is 72,928 km2. Although it is not qualified with astatus of threatened species, we estimated a decrease of almost20% of its habitat area in a relatively short time, particularlyin the northern populations, associated with Mediterranean-influenced climate, which is at least worrisome consideringwe used the most conservative future climate scenario. Fusset al. [31] showed that the actual reality is the worst climatescenario projected. The 𝐾-means analyses, applied to fourbest least correlated variables of WorldClim (BIO2, BIO4,BIO12, and BIO15), identified seven strata groups of N.dombeyi in Chile at very broad spatial scales (Figure 1).Thesestrata represented relatively homogeneous habitats accordingto the climatic key variables mentioned. The strata namedSeptentrional (#1), Alto Biobı́o (#2), Los Lagos Andes (#7),and Patagonia (#3) form the latitudinal gradient across AndesMountain range including higher altitudes, while the strataNahuelbuta (#4), Araucanı́a y Los Rı́os (#5), and Los LagosCosta (#6) present some oceanic influence at a gradientlocated in Coastal Mountain range and central valley of thecentral distribution of N. dombeyi at lower elevation. Thestrata were coherently identified as the bioclimatic transitionfrom temperate biome with some Mediterranean influence,especially in the northern portion of the Septentrional stra-tum and then the typical temperate bioclimatic classificationacross the rest of the distribution of the species [32].

    3.2. Assessment of N. dombeyi Genotypes Sampled by GBSUsing ApeKI. Weobtainedmore than 17Gb out of 172,171,356reads, from which 85.82% keep a score 𝑄 ≥ 30. From the96 barcoded samples in our study, an average of 1.5 million(SD 1.7 million) sequence reads per sample were obtained(range from 278 to 8.2 millions) (see Supplementary Figure2). After removal of SNP with a minor allele frequency0.05 and removing samples and SNP with more than 90%“NN” (unassigned) genotypes, the dataset consisted of 73

  • International Journal of Genomics 5

    0 100 200 300 40050

    Ecological strataof N. dombeyi:Septentrional:

    Alto Biobío:

    Patagonia:

    Nahuelbuta:

    Araucanía y Los Ríos:

    Los Lagos Costa:

    Los Lagos Andes:1

    2

    3

    4

    5

    6

    7

    48∘S

    47∘S

    46∘S

    45∘S

    44∘S

    43∘S

    42∘S

    41∘S

    40∘S

    39∘S

    38∘S

    37∘S

    36∘S

    35∘S

    34∘S

    48∘S

    47∘S

    46∘S

    45∘S

    44∘S

    43∘S

    42∘S

    41∘S

    40∘S

    39∘S

    38∘S

    37∘S

    36∘S

    35∘S

    34∘S

    77∘W 75∘W 71∘W73∘W 72∘W 70∘W 67∘W69∘W 68∘W74∘W76∘W

    77∘W

    N

    75∘W 71∘W73∘W 72∘W 70∘W 67∘W69∘W 68∘W74∘W76∘W

    (km)

    Figure 1: Using ecological niche models to identify different strata for Nothofagus dombeyi in Chile.

    individuals with 10,109 binary SNP from seven strata. Thisdataset was subjected to the impute step and finally totalnumbers of “NN” genotypes were 19,067 (0.03%) missingdata.

    3.3. Putative under Selection SNP (Adaptive SNP). Usingthe LOSITAN approach, 124 adaptive SNP (1.2%) wereidentified (directionally selected) at the false discovery rate(FDR) threshold of 0.1 (Table 2; Supplementary Table 1). Wedetected 99 adaptive SNP in BayeScan and there were 48overlap cases between the loci reported by each method(see Supplementary Table 1). The adaptive SNP detectedby both approaches were limited (39%), possibly reflecting

    discrepancies in their methodologies. However, all the 124SNP identified by LOSITAN had Fst values ≥ 0.30 (estimatedacross the 73 N. dombeyi individuals and 10,109 loci), whichmay be considered as strongly differentiated.

    Using Sam𝛽ada we detected 2,406 significant genotypesassociated with a given environmental variable. They rep-resent 884 SNP (8.7% of total SNP assessed) and only3% correspond to heterozygous genotypes. From the 124outlier SNP identified by LOSITAN, genotypes in 121 SNPwere identified as associated with environmental variablesby Sam𝛽ada (see Supplementary Table 1). Genetic differen-tiation associated with both temperature and precipitationgradients was detected.

  • 6 International Journal of Genomics

    Table 2: Representative list of outlier single nucleotide polymorphisms (SNP) as putative candidates for adaptation in Nothofagus dombeyiin Chile and their significant associations with environmental variables using Sam𝛽ada. Italic values show an SNP that is not a significantoutlier.

    Locus SNP IDLOSITAN BayeScan Sam𝛽ada

    Het Fst 𝑃(Simul Fst < sample Fst) BIO2 BIO4 BIO5 BIO6 BIO12 BIO15 BIO18 BIO19

    TP3479 4 0.368 0.545 0.99999 Yes x x x xTP8429 14 0.282 0.496 0.99996 — x x x xTP23119 — 0.491 0.035 0.46921TP30254 42 0.385 0.442 0.99996 Yes x x x xTP30852 43 0.438 0.438 1.00000 Yes x x x xTP32296 44 0.467 0.377 0.99991 Yes x x x xTP38613 50 0.364 0.618 1.00000 Yes x x x xTP65805 91 0.540 0.488 1.00000 Yes x x

    Hs Ht Dst Htp Dstp Fst Fstp Fis DestHo

    −1.0

    −0.5

    0.0

    0.5

    1.0

    (a)

    Hs Ht Dst Htp Dstp Fst Fstp Fis DestHo

    −0.2

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    (b)

    Figure 2: Summary statistics of genetic variation existing inNothofagus dombeyi in Chile estimated by 10,109 neutral SNP (a) or 124 adaptiveSNP (b). Ho: heterozygosity within population; Hs: genetic diversity within population; Ht: overall gene diversity; Http: corrected Ht; Dst:gene diversity among samples; Dstp: corrected Dst; Fst: fixation index; Fstp: corrected Fst; Fis: inbreeding coefficient per overall loci; Dest:measure of population differentiation.

    3.4. Genome-Wide Genetic Variation and Differentiation.Using neutral plus outlier SNP (10,109), different basic statis-tics of genetic variation of N. dombeyi in Chile show lowto medium genetic diversity level and low level of stratadifferentiation (Figure 2). The low genetic structure foundindicates relatively high gene flow, which is consistent withthe fact that N. dombeyi has a distribution more or lesscontinuous and is wind pollinated. This result is similar toother results obtained with the same and other species ofNothofagus gender [33, 34]. With outlier SNP, the indicatorsbased on heterozygosity increase and 𝐹 statistics shows cleardifferences within strata.

    Overall average pairwise Fst values among populations,calculated over the set of the 10,109 SNP, show medium pop-ulation structuring across the distribution of N. dombeyi inChile (Table 3). The average pairwise Fst values ranged from0.062 between two very close strata (4 and 5) in the Coastal

    Table 3: Average pairwise Fst among seven strata with Nothofagusdombeyi presence in Chile based on a set of 10,109 single nucleotidepolymorphisms (SNP). Fst 0.15–0.2 are moderately differentiated inbold font and Fst > 0.25 are considered strongly differentiated inbold and italic font.

    Strata 1 2 3 4 5 62 0.0743 0.258 0.1934 0.111 0.074 0.1835 0.109 0.066 0.138 0.0626 0.156 0.108 0.140 0.096 0.0697 0.172 0.121 0.141 0.110 0.081 0.096

    Mountains to 0.258 between strata 1 and 3, both in AndeanMountains but in the extremes of distribution (Table 1).

  • International Journal of Genomics 7

    Cluster 4Cluster 1Cluster 2Cluster 3

    Cluster 5

    20 40 600PCA axis

    020406080

    Cum

    ulat

    edva

    rianc

    e (%

    )

    Figure 3: Discriminant Analysis of Principal Components (DAPC)scatterplot drawnusing 124 outlier single nucleotide polymorphisms(SNP) across 73 Nothofagus dombeyi individuals in the R packageadegenet. Dots represent individuals, with colours denoting clusterallocation. Percentages of cumulated variance explained byPrincipalComponent 1 (PC1) to PC20 are shown in the bottom left corner.Minimum spanning tree based on the (squared) distances betweenclusters within the entire space is shown.

    The Principal Coordinates Analysis (PCoA) plot revealed ageographically ordered pattern (see Supplementary Figure 3).The first PC suggests the existence of two clades in the data,while the second one shows groups of closely related isolatesarranged along a cline of genetic differentiation. Premoli [35]found continuous clinal genetic variation in populations ofN.pumilio along the altitudinal gradient, as a result of adaptiveresponses to ecological gradients and/or restrictions for geneflow.The same pattern is shown by neutral SNP and adaptiveSNP but with a more clear delimitation in the latter case.This structure was confirmed by a neighbour-joining (NJ)tree (Supplementary Figure 4). Again higher resolution isachieved with adaptive SNP. As expected, both approachesgive congruent results, but both are complementary; NJshows bunches of related individuals, but the cline of geneticdifferentiation is much clearer in PCA.

    Visualisation of broad-scale population structure using aDAPC (Figure 3 and Table 4) with 124 outlier SNP revealedtwo distinct genetic groups. One group included individualssampled in all strata corresponding to Clusters 1, 2, 3, and 5(Table 4), and the other group comprised only ten individualsof stratum 3 plus one individual of stratum 2 forming Cluster4 (represented almost exclusively by individuals of Patagoniastratum). Also the actual proximities between clusters showthe great genetic distance of Cluster 4 with the rest.This resultis consistent with a north–south phylogenetic divergence at

    Table 4: Individual allocation to five genetic clusters accordingto strata determined by niche modelling of N. dombeyi in Chile.Clusters were identified by find.cluster function of adegenet Rpackage using 124 adaptive single nucleotide polymorphisms (SNP).

    Strata Cluster assignationC1 C2 C3 C4 C5

    1 — — 8 — —2 — — 12 1 —3 — — — 10 44 8 — 1 — —5 10 5 1 — —6 — 7 — — —7 — 6 — — —

    c. 43∘S found within subgenus Nothofagus in southern SouthAmerica, including N. dombeyi [12, 36].

    Clearly, when examining patterns of N. dombeyi popula-tion differentiation at neutral versus adaptive SNP, we havedetected several distinct differences. Primarily stochasticprocesses drive differentiation of neutral SNP, whereas bothselective and stochastic processes drive that of adaptive SNP,for example, [37–39].This result supports the hypothesis thatputative under selection SNPare being affected by adaptation,which is not affecting the neutral SNP. The next step is tocharacterize the phenotypic outcomes of alternative geno-types to confirm if themechanism is genetic drift or selection,similar to previous studies in Nothofagus pumilio [40, 41].Ideally, phenotype-genotype association studies should befollowed by the profiling of gene expression, functional tests,and selection tests to determine that a gene or genes areinvolved in shaping an adaptive trait [42].

    3.5. Loci Contributions. Over 24% of the adaptive SNPdetected by LOSITAN had a high loading (here defined as≥0.02) in one or two of the three PC in a PCA with morethan 10K loci of the Chilean population of N. dombeyi(Supplementary Figures 5, 6, and 7), thus confirming itscontribution to population structure. Adaptive SNP ID 4,14, 42, 50, and 91 are five of the most discriminating lociand their allele frequencies over seven strata show a highor even dominant allele frequency in stratum 3 (Figure 4).This is the southernmost stratum (#3 Patagonia) and thismayjustify the differences in allele frequencies. Irrespective ofthe mechanism underlying these changes (drift or selection),this illustrates that, in the natural distribution of N. dombeyiin Chile, specific nucleotides can undergo drastic changeswithin only a hundred kilometres of distance. Our resultsexhibited spatial patterns differentiating one stratum in lociwhich could occur in genomic regions or genes for importantfunctions of N. dombeyi across their habitat. For example,the GBS sequence tags of adaptive SNP ID 43 (TP30852)show a high identity with a heat shock protein or adaptiveSNP ID 44 (TP32296) with histone acetyltransferase enzyme(data not shown). The first is fundamental for heat orother environmental stresses and the second can be involvedin drought sensing or another response to environmental

  • 8 International Journal of Genomics

    tt

    t

    t tt

    t

    SNP 4 #TP3479

    aa

    a

    a aa

    a

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Alle

    le fr

    eque

    ncy

    3 5 71Strata

    cc

    c

    c

    c

    c

    c

    SNP 14 #TP8429

    tt

    t

    tt

    t

    t

    3 5 71Strata

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Alle

    le fr

    eque

    ncy

    aa

    a

    a

    a a

    a

    SNP 42 #TP30254

    t t

    t

    t

    t t

    t

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Alle

    le fr

    eque

    ncy

    3 5 71Strata

    tt

    t

    t

    tt

    t

    SNP 50 #TP38613

    cc

    c

    c

    cc

    c

    3 5 71Strata

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    Alle

    le fr

    eque

    ncy

    gg g

    g

    g g

    g

    SNP 91 #TP65805

    aa a

    a

    a a

    a

    0.2

    0.4

    0.6

    0.8

    Alle

    le fr

    eque

    ncy

    3 5 71Strata

    a

    aa

    a a

    a a

    SNP #TP23119

    t

    t t

    tt

    tt

    3 5 71Strata

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    Alle

    le fr

    eque

    ncy

    Figure 4: Spatial distributions of single nucleotide polymorphism (SNP) loci/genotypes in N. dombeyi in Chile for each strata. The graphsshow frequencies of loci/genotypes differentiating among strata (SNP #3479, SNP #8429, SNP #30254, SNP #38613, and SNP #65805) and nodifferentiating among strata (SNP #23119).

    factors. We therefore suggest that these two SNP may beinteresting candidates for future functional studies thatmightfacilitate other studies, for example, [43] focused on thedevelopment of genomic resources for Nothofagus species.However, improvements in study design and analyses ofreplicated studies will be needed before this very promisingapproach can be brought to application for managing geneticresources [44].

    4. Conclusions

    Development of 10,109 genome-wide SNP for N. dombeyiusing GBS made evaluation of genomic diversity and fine-scale population structure possible for the first time in thisspecies in Chile. Results showed that genome-wide patternsof genetic diversity and differentiation varied widely acrossthe genome. As such, we identified numerous genomic

    regions exhibiting signatures of divergent selection. We havealso provided strong evidence of substantial genetic differ-entiation associated with both temperature and precipitationgradients, suggesting local adaptation.

    Competing Interests

    The authors declare that there are no competing interestsregarding the publication of this paper.

    Acknowledgments

    The authors thank CONAF (Corporación Nacional Fore-stal) and SNASPE (Sistema Nacional de Áreas SilvestresProtegidas del Estado) for providing permission to sampleplant material and especially thank José Luis Pérez for

  • International Journal of Genomics 9

    sampling stratum 3 (Patagonia). Angela Sierra, Daniela Fer-nandez, Giannina Espinoza, and Joseline Valdés are grate-fully acknowledged for providing technical and scientificassistance. The authors thank the University of WisconsinBiotechnology Center DNA Sequencing Facility for provid-ing sequencing and support services. Funding for this projectwas provided by the Research Fund for Native Forest ofCONAF (Project 068/2012).

    References

    [1] R. Frankham, J. D. Ballou, and D. A. Briscoe, A Primer of Con-servation Genetics, Cambridge University Press, Cambridge,UK, 2004.

    [2] R. Frankham, “Where are we in conservation genetics andwhere do we need to go?” Conservation Genetics, vol. 11, no. 2,pp. 661–663, 2010.

    [3] D. J. Fraser and L. Bernatchez, “Adaptive evolutionary conserva-tion: towards a unified concept for defining conservation units,”Molecular Ecology, vol. 10, no. 12, pp. 2741–2752, 2001.

    [4] B. Gebremedhin, G. F. Ficetola, S. Naderi et al., “Frontiersin identifying conservation units: from neutral markers toadaptive genetic variation,” Animal Conservation, vol. 12, no. 2,pp. 107–109, 2009.

    [5] F. W. Allendorf, P. A. Hohenlohe, and G. Luikart, “Genomicsand the future of conservation genetics,”Nature Reviews Genet-ics, vol. 11, no. 10, pp. 697–709, 2010.

    [6] C. Y. Tang, C.-L. Hung, H. Zheng, C.-Y. Lin, and H.Jiang, “Novel computational technologies for next-generationsequencing data analysis and their applications,” InternationalJournal of Genomics, vol. 2015, Article ID 254685, 3 pages, 2015.

    [7] B. B. Patnaik, S. Y. Park, S.W. Kang et al., “Transcriptome profileof the asian giant hornet (Vespa mandarinia) using illuminaHiSeq 4000 sequencing: De novo assembly, functional anno-tation, and discovery of SSR markers,” International Journal ofGenomics, vol. 2016, Article ID 4169587, 15 pages, 2016.

    [8] A. Bonin, F. Nicole, F. Pompanon, C. Miaud, and P. Taberlet,“Population adaptive index: a new method to help measureintraspecific genetic diversity and prioritize populations forconservation,” Conservation Biology, vol. 21, no. 3, pp. 697–708,2007.

    [9] D. Demontis, C. Pertoldi, V. Loeschcke, K. Mikkelsen, T. Axels-son, and T. N. Kristensen, “Efficiency of selection, as measuredby single nucleotide polymorphism variation, is dependent oninbreeding rate inDrosophila melanogaster,”Molecular Ecology,vol. 18, no. 22, pp. 4551–4563, 2009.

    [10] H. Kirk and J. R. Freeland, “Applications and implicationsof neutral versus non-neutral markers in molecular ecology,”International Journal of Molecular Sciences, vol. 12, no. 6, pp.3966–3988, 2011.

    [11] A. C. Premoli, M. P. Quiroga, P. Mathiasen, and T. Kitzberger,“Ecological niche modeling meets phylogeography to unravelhidden past history of key forest genera in plant geography:Podocarpus and Nothofagus,” Natureza & Conservação, vol. 10,no. 2, pp. 160–168, 2012.

    [12] M. C. Acosta, P. Mathiasen, and A. C. Premoli, “Retracing theevolutionary history of Nothofagus in its geo-climatic context:new developments in the emerging field of phylogeology,”Geobiology, vol. 12, no. 6, pp. 497–510, 2014.

    [13] A. C. Premoli, P. Mathiasen, M. Cristina Acosta, and V.A. Ramos, “Phylogeographically concordant chloroplast DNA

    divergence in sympatric Nothofagus s.s. How deep can it be?”New Phytologist, vol. 193, no. 1, pp. 261–275, 2012.

    [14] D. Alarcón and L. A. Cavieres, “In the right place at the righttime: habitat representation in protected areas of South Amer-ican Nothofagus-dominated plants after a dispersal constrainedclimate change scenario,” PLoS ONE, vol. 10, no. 3, Article IDe0119952, 2015.

    [15] W. Thuiller, B. Lafourcade, R. Engler, and M. B. Araújo,“BIOMOD—a platform for ensemble forecasting of speciesdistributions,” Ecography, vol. 32, no. 3, pp. 369–373, 2009.

    [16] R. J. Hijmans, S. E. Cameron, J. L. Parra, P. G. Jones, and A.Jarvis, “Very high resolution interpolated climate surfaces forglobal land areas,” International Journal of Climatology, vol. 25,no. 15, pp. 1965–1978, 2005.

    [17] T. Calinski and J. Harabasz, “A dendrite method for clusteranalysis,” Communications in Statistics, vol. 3, pp. 1–27, 1974.

    [18] R. J. Elshire, J. C. Glaubitz, Q. Sun et al., “A robust, simplegenotyping-by-sequencing (GBS) approach for high diversityspecies,” PLoS ONE, vol. 6, no. 5, Article ID e19379, 2011.

    [19] F. Lu, A. E. Lipka, J. Glaubitz et al., “Switchgrass genomicdiversity, ploidy, and evolution: novel insights from a network-based SNP discovery protocol,” PLoS Genetics, vol. 9, no. 1,Article ID e1003215, 2013.

    [20] J. C. Glaubitz, T. M. Casstevens, F. Lu et al., “TASSEL-GBS: ahigh capacity genotyping by sequencing analysis pipeline,”PLoSONE, vol. 9, no. 2, article e90346, 2014.

    [21] D. Money, K. Gardner, Z. Migicovsky, H. Schwaninger, G.-Y.Zhong, and S. Myles, “LinkImpute: fast and accurate genotypeimputation for nonmodel organisms,” G3: Genes, Genomes,Genetics, vol. 5, no. 11, pp. 2383–2390, 2015.

    [22] T. Antao, A. Lopes, R. J. Lopes, A. Beja-Pereira, and G. Luikart,“LOSITAN: a workbench to detect molecular adaptation basedon a Fst-outliermethod,”BMCBioinformatics, vol. 9, article 323,2008.

    [23] M. Foll and O. Gaggiotti, “A genome-scan method to identifyselected loci appropriate for both dominant and codominantmarkers: a Bayesian perspective,” Genetics, vol. 180, no. 2, pp.977–993, 2008.

    [24] S. Joost, A. Bonin, M. W. Bruford et al., “A spatial analysismethod (SAM) to detect candidate loci for selection: towardsa landscape genomics approach to adaptation,”Molecular Ecol-ogy, vol. 16, no. 18, pp. 3955–3969, 2007.

    [25] A. J. Dobson and A. G. Barnett, An Introduction to GeneralizedLinear Models, Chapman and Hall/CRC, London, UK, 3rdedition, 2008.

    [26] T. Jombart, “adegenet: a R package for the multivariate analysisof geneticmarkers,”Bioinformatics, vol. 24, no. 11, pp. 1403–1405,2008.

    [27] T. Jombart and I. Ahmed, “adegenet 1.3-1: new tools for theanalysis of genome-wide SNP data,” Bioinformatics, vol. 27, no.21, Article ID btr521, pp. 3070–3071, 2011.

    [28] J. Goudet, “HIERFSTAT, a package for R to compute and testhierarchical F-statistics,” Molecular Ecology Notes, vol. 5, no. 1,pp. 184–186, 2005.

    [29] M. Nei, Molecular Evolutionary Genetics, Columbia UniversityPress, New York, NY, USA, 1987.

    [30] L. Jost, “GST and its relatives do not measure differentiation,”Molecular Ecology, vol. 17, no. 18, pp. 4015–4026, 2008.

    [31] S. Fuss, J. G. Canadell, G. P. Peters et al., “Betting on negativeemissions,” Nature Climate Change, vol. 4, no. 10, pp. 850–853,2014.

  • 10 International Journal of Genomics

    [32] F. Luebert and P. Pliscoff, Sinopsis Bioclimática y Vegetacional deChile, Editorial Universitaria, Santiago, Chile, 2006.

    [33] A. C. Premoli, “Genetic variation in a geographically restrictedand two widespread species of South American Nothofagus,”Journal of Biogeography, vol. 24, no. 6, pp. 883–892, 1997.

    [34] C. J. Duncan, J. R. P. Worth, G. J. Jordan, R. C. Jones, and R.E. Vaillancourt, “Genetic differentiation in spite of high geneflow in the dominant rainforest tree of southeastern Australia,Nothofagus cunninghamii,” Heredity, vol. 116, no. 1, pp. 99–106,2016.

    [35] A. C. Premoli, “Isozyme polymorphisms provide evidence ofclinal variation with elevation in Nothofagus pumilio,” Journalof Heredity, vol. 94, no. 3, pp. 218–226, 2003.

    [36] A. C. Premoli, P. Mathiasen, M. C. Acosta, and V. A. Ramos,“Phylogeographically concordant chloroplast DNA divergencein sympatric Nothofagus s.s. How deep can it be?” New Phytol-ogist, vol. 193, no. 1, pp. 261–275, 2012.

    [37] J. Galindo, P. Morán, and E. Rolán-Alvarez, “Comparing geo-graphical genetic differentiation between candidate and non-candidate loci for adaptation strengthens support for parallelecological divergence in the marine snail Littorina saxatilis,”Molecular Ecology, vol. 18, no. 5, pp. 919–930, 2009.

    [38] A. V. Stronen, B. Jedrzejewska, C. Pertoldi et al., “Genome-wide analyses suggest parallel selection for universal traitsmay eclipse local environmental selection in a highly mobilecarnivore,” Ecology and Evolution, vol. 5, no. 19, pp. 4410–4425,2015.

    [39] B. Guo, J. DeFaveri, G. Sotelo, A.Nair, and J.Merilä, “Populationgenomic evidence for adaptive differentiation in Baltic Seathree-spined sticklebacks,” BMC Biology, vol. 13, article 19, 2015.

    [40] A. C. Premoli and C. A. Brewer, “Environmental v. geneticallydriven variation in ecophysiological traits ofNothofagus pumiliofrom contrasting elevations,” Australian Journal of Botany, vol.55, no. 6, pp. 585–591, 2007.

    [41] A. C. Premoli, E. Raffaele, and P. Mathiasen, “Morphologicaland phenological differences in Nothofagus pumilio from con-trasting elevations: evidence from a common garden,” AustralEcology, vol. 32, no. 5, pp. 515–523, 2007.

    [42] C. Pardo-Diaz, C. Salazar, and C. D. Jiggins, “Towards theidentification of the loci of adaptive evolution,” Methods inEcology and Evolution, vol. 6, no. 4, pp. 445–464, 2015.

    [43] V. A. El Mujtar, L. A. Gallo, T. Lang, and P. Garnier-Géré,“Development of g