苹果intoy一docker 停止运行容器

正在初始化报价器苹果手机电不耐用怎么办3个回答路过狂艹本吧32iphone电池不耐用,让电池更耐用必知技巧:
  第一、保持正常的使用频率
  第一件事是你要确保你不会整天抱着你的iPhone,iPad或者iPod。这听起来似乎很蠢,但是当你的设备有了新的功能,或者你有了全新的iPhone你肯定会舍不得放手,因此你的电池可能会很快被耗尽。升级到IOS6以后,会有比之前版本更多的通知,定位以及其他消耗电量的功能,而iPhone5拥有一块更大的屏幕和LTE网络会耗费更多的电量。
  第二、操作系统和设备是否有问题
  如果你电池续航一直都很差而且你能看到你的电量一直的不断地减少,这里有一些建议给你试试。
  1、重新设备或者恢复你的设备。如果你很长一段时间没有重启过你的设备,请重启。因为一些恶意进程和程序可能会做一些他们不该做的事情,而重启可以修复这个问题。
  2、电池周期
  大约一个月一次,如果你感觉到你的电池有问题有应该彻底用完你的电池电量,一直用到关机为止,然后充电到满。
  3、恢复你的设备到出厂状态。
  导致IOS设备电池续航问题的最大原因是当他们恢复备份时并没有恢复出厂设置。虽然很繁琐,虽然你可能会损失一些东西,但他可能是解决电池续航问题的最方法。这是最好的选择,你需要重新设置之前个人喜好选项,你将失去之前的数据包括游戏存档等等。但是恢复过后在正常使用情况下你的电池寿命肯定会比以前好。
  4、去Apple store。
  有时候电池续航能力不行iPhone,iPad或者iPod是出厂时就存在的先天问题。你只有前去Apple store修理或者更换设备才能解决。
  第三、为iPhone补充电量
  不要不好意思为你的iPhone补充电量,如果你真的需要长时间使用你的iPhone,iPad或者iPod,你就要尽可能抓住每一个机会为你的IOS设备补充电量。在家,在公司,在汽车里等等有很多机会你可以为你的iPhone补充电量。
  当然这对于iPhone5来说稍微有些棘手,因为他使用了新的接口,在各种环境下充电你可能需要价值不菲的转接头。但是如果你真的需要这些钱也是值得的。
  第四、关掉不使用的进程
  任何在你的iPhone,iPad,iPod上运行的进程都会消耗电量。为了保证你的续航时间,你可以选择性的关掉一些不需要的功能,所以你需要在使用和电量中间寻找一个平衡点。尽可能的满足自己的使用需求之后,在尽可能的减少不必要的功能和进程。
  1、关闭Siri的Raise to speak功能。进入“设置”→“通用”→“Siri”。许多朋友反映关掉这个功能可以节省出不少的电量。
  2、关闭定位服务。进入“设置”→“隐私”→“定位服务”关闭不需要使用定位功能的APP。
  3、关闭推送通知。同样地进入“设置”→“通知”。关闭那些你不需要推送消息的APP。
  4、双击Home键可以看到最近使用的APP程序,用手指按住一个APP图标进入“摇晃”状态并且关闭可能正在运行的运用程序,特别是网络电话如skype。
  一些其他省电设置:
  设置自动锁定时间为1分钟
  关闭任何其他的声音,比如键盘点击声音
  关闭iPod EQ
  听音乐和音频时使用耳机
  调低屏幕亮度
  不使用时关闭蓝牙
  不使用时关闭Wi-Fi
  不使用时关闭LTE
  关闭蜂窝应用程序和媒体下载
  关闭所有电子邮件,日历,联系人账号的的推送功能
挥手说seeyou1.在不用设备时,可以通过按下电源键切换到休眠模式。
2.开启省电模式:设定-(一般)-省电模式-滑动开启。
3.通过任务管理器(长按HOME键两秒)关闭不必要的应用程序。
4.不影响使用的情况下关闭蓝牙,定位,推送通知等设置以延长待机时间。
5.取消应用程序的自动同步功能。
6.减少背景灯时间。
7.调低显示屏的亮度。
8.有条件的话更换手机原厂电池尝试。
一龙回家你好,1,使用非原装充电器或者移动电源为手机充电,虽然电量显示已满,但很可能是虚电,可以用原装充电器来回充放几次,即可恢复正常。
2,后台程序运行过多,消耗大量电量,苹果的后台程序优化没有想象的那么好,建议将没用的后台程序关闭再进行尝试。
3,电池损坏,在换电池之前可以尝试刷机,没有用的话建议更换电池。
热门问答123456789101112131415161718192021222324252627282930相关问答3个回答非法组合首先检查一下插充电器的插座有没有打开,有没有电源,这是最基本的,第一是有电源,有的人会很笨,自己的手机充电的充电线的插头都插在没有电的茶盘上。解决办法:插座一定要有电。要不是没开电...1个回答奈落00375你好,1新电池头三次先用至低电告警,再配合原装直充在手机开机充电到满,然后继续保持充电约1小时。2日常使用中,电池充满就好,要避免继续长时间充电,建议用到告警就好,万一用到关机,电...4个回答狄梦华你好! 主板有漏电的结果。掉水里后还没有把水份全部排干,最好先不要用,放入有干燥剂的密封容器内干燥或放在小功率的灯泡下烘干。希望对你有帮助!
3个回答esf-seco1、手机系统故障,因此导致电池电量显示出现问题。可以直接关机重启。
2、充电器出现故障。换上别的充电器试一下。
3、充电器和手机接触不良。充电器没有插好。解决办法是来回换插...3个回答梦想星期一都为锂电池;
  “锂电池”,是一类由锂金属或锂合金为负极材料、使用非水电解质溶液的电池。1912年锂金属电池最早由Gilbert N. Lewis提出并研究。20世纪70年...3个回答西红柿好IP17908主叫绑定业务是通过17908接入号,根据语音提示拨打区号+对方电话号码(亦可拨卡号和密码使用),拨打国内和国际长途电话。从而提供您先用后付的IP电话服务,每月出帐一次...3个回答___筱莜丶084苹果手机费电,省电方法如下
1、苹果手机耗电量很快,有可能手机一直在使用、手机后台运行程序较多、手机上面有很多缓存文件没有清理,所以导致手机在运行的时候需要夹杂的数据较多,中...3个回答TA000CEbgd1,点击打开设置
2,下拉页面,找到电话
3,点击打开,一眼即可看到呼叫转移,点击打开
4,载入呼叫转移,需要5S左右
5,呼叫转移开启,有全部通话和自定义...6个回答青仔9315关闭后台应用程序刷新,关闭电子邮件推送,关闭应用程序推送通知,关闭自动下载,降低屏幕的亮度,关掉振动提示,缩短自动锁屏时间,关掉AirDrop,停止定位服务,关掉蓝牙,关掉不必要的...3个回答JRH3277充电宝基本是每个手机都是适用的,有的充电宝还有多个USB街头,充电宝使用时间的长短,首先是看你的移动电源的质量,其次是你对移动电源保养。在充电宝电池的上方一般会有三个插口,主要有充...欢迎来到慢慢买! &
扫描二维码下载客户端
慢慢买客户端
苹果(Apple) MacBook MNYJ2CH/A 12英寸笔记本电脑 I5处理器/8G/512GSSD/集显
价格区间:¥11264至¥12288
优选评价:4.6分
全网共有 5321 人评论 &
& 类目排名第
品牌:上市时间:2017年06月屏幕尺寸:12英寸CPU类型:intel i5产品类型:娱乐办公本显存容量:核心显卡/集显
苹果 MNYJ2CH/A怎么样?看看下面购买过的网友怎么说!
49人参与评分
购买过此商品?
电脑很轻薄,还是挺好用的,五星吧
购买自【】
电脑配送不错,速度快,电脑没有想象中的那么好,运行速度尚可,但是很轻便,日常办公够用了,就是耗电貌似快点似的
购买自【】
最低¥7499 最低¥7899 最低¥8999 最低¥4599 最低¥5299 最低¥6999 最低¥6499 最低¥5799 最低¥6299 最低¥9988 最低¥10399 最低¥8999
购买过苹果 MNYJ2CH/A吗?快来说说苹果 MNYJ2CH/A怎么样,给更多网友参考吧!
千万不要买
&11264&11599&12188&12288&12288
笔记本电脑口碑排行
最低价:¥6299
最低价:¥6199
最低价:¥6599
最低价:¥7099
最低价:¥8499
笔记本电脑购买知识
扫描下载慢慢买APP
你的随身购物助手
慢慢买是一个倡导理性消费的导购平台,7年来我们专注为用户推荐高性价比的商品,同时开发了全网比价、历史价格查询等购物助手,力求帮助消费者花更少的钱买到优质的商品。
合作商城:> Yoddish Translator
请选择您要上传的头像图片,完成后关闭此弹框即可。
&&&&您所提交的评论已被保存,登录后即可发表
&&&&您已经选择过了。
应用大小:6.9 MB
类别:娱乐
当前版本:1.0
语言:英语
运行环境:需要 iOS 8.0 或更高版本。与 iPhone、iPad、iPod touch 兼容。
下载此APP的用户也下载
简介:Yoddish translator rearranges sentences into...
Yoddish Translator应用说明
Yoddish translator rearranges sentences into OSV (Object Subject Verb) structure. Just type in normal sentences in English and Yoddish translator will rearrange it in OSV structure and send it back to you! Very fun and educational tool for children of all ages.An example of OSV word order would be: "Very powerful, master Yoda is."Which in a normal English (SVO structure): "Master Yoda is very powerful".
| | | | | | |
| | | | | | | |
|||||||||||||||
iTunes分类排行榜
||||||||||||||||||
京公网安备98
苹果园为iOS用户提供和下载,最新的、、、等,分享最权威的资讯、、及解决办法,拥有最火爆的,苹果园一家专注解决iOS所求的网站。
05:36:21.849---- 05:36:21.865----15.6Loading metrics
Open Access
Peer-reviewed
Research Article
New Insight into the History of Domesticated Apple: Secondary Contribution of the European Wild Apple to the Genome of Cultivated Varieties
New Insight into the History of Domesticated Apple: Secondary Contribution of the European Wild Apple to the Genome of Cultivated Varieties
Amandine Cornille,&
Pierre Gladieux,&
Marinus J. M. Smulders,&
Isabel Roldán-Ruiz,&
Fran?ois Laurens,&
Bruno Le Cam,&
Anush Nersesyan,&
Joanne Clavel,&
Marina Olonova,&
Laurence Feugey
AbstractThe apple is the most common and culturally important fruit crop of temperate areas. The elucidation of its origin and domestication history is therefore of great interest. The wild Central Asian species Malus sieversii has previously been identified as the main contributor to the genome of the cultivated apple (Malus domestica), on the basis of morphological, molecular, and historical evidence. The possible contribution of other wild species present along the Silk Route running from Asia to Western Europe remains a matter of debate, particularly with respect to the contribution of the European wild apple. We used microsatellite markers and an unprecedented large sampling of five Malus species throughout Eurasia (839 accessions from China to Spain) to show that multiple species have contributed to the genetic makeup of domesticated apples. The wild European crabapple M. sylvestris, in particular, was a major secondary contributor. Bidirectional gene flow between the domesticated apple and the European crabapple resulted in the current M. domestica being genetically more closely related to this species than to its Central Asian progenitor, M. sieversii. We found no evidence of a domestication bottleneck or clonal population structure in apples, despite the use of vegetative propagation by grafting. We show that the evolution of domesticated apples occurred over a long time period and involved more than one wild species. Our results support the view that self-incompatibility, a long lifespan, and cultural practices such as selection from open-pollinated seeds have facilitated introgression from wild relatives and the maintenance of genetic variation during domestication. This combination of processes may account for the diversification of several long-lived perennial crops, yielding domestication patterns different from those observed for annual species.
Author Summary
The apple, one of the most ubiquitous and culturally important temperate fruit crops, provides us with a unique opportunity to study the process of domestication in trees. The number and identity of the progenitors of the domesticated apple and the erosion of genetic diversity associated with the domestication process remain debated. The Central Asian wild apple has been identified as the main progenitor, but other closely related species along the Silk Route running from Asia to Western Europe may have contributed to the genome of the domesticated crop. Using rapidly evolving genetic markers to make inferences about the recent evolutionary history of the domesticated apple, we found that the European crabapple has made an unexpectedly large contribution to the genome of the domesticated apple. Bidirectional gene flow between the domesticated apple and the European crabapple resulted in the domesticated apple being currently more similar genetically to this secondary genepool than to the ancestral progenitor, the Central Asian wild apple. We found that domesticated apples have evolved over long time scales, with contributions from at least two wild species in different geographic areas, with no significant erosion of genetic diversity. This process of domestication and diversification may be common to other fruit trees and contrasts with the models documented for annual crops.
IntroductionDomestication is a process of increasing codependence between plants and animals on the one hand, and human societies on the other , . The key questions relating to the evolutionary processes underlying domestication concern the identity and geographic origin of the wild progenitors of domesticated species , the nature of the genetic changes underlying domestication , , the tempo and mode of domestication (e.g., rapid transition versus protracted domestication)
and the consequences of domestication for the genetic diversity of the domesticated species , , , . An understanding of the domestication process provides insight into the general mechanisms of adaptation and the history of human civilization, but can also guide modern breeding programs aiming to improve crops or livestock species further , .
Plant domestication has mostly been studied in seed-propagated annual crops, in which strong domestication bottlenecks have often been inferred, especially in selfing annuals, such as foxtail millet, wheat and barley , , , , , . Genetic data have suggested that domestication or the spread of domesticated traits has been fairly rapid in some annual species (e.g, maize or sunflower), with limited numbers of populations or species contributing to current diversity , , , , , . In contrast, a combination of genetics and archaeology suggested a protracted model of domestication for other annual crops, and in particular for the origin of wheat or barley in the Fertile Crescent , . However, the genetic consequences of domestication have been little investigated in long-lived perennials, such as fruit trees , , . Trees have several biological features that make them fascinating and original models for investigating domestication: they are outcrossers with a long lifespan and a long juvenile phase, and tree populations are often large and connected by high levels of gene flow , .
Differences in life-history traits probably result in marked differences in the mode and speed of evolution between trees and seed-propagated selfing annuals , , . For example, outcrossing may tend to make domestication more difficult, in part because the probability of fixing selected alleles is lower than in selfing crops , . The combination of self-incompatibility and a long juvenile phase also results in highly variable progenies, making breeding a slow and expensive process, and rendering crop improvement difficult. The development of vegetative propagation based on cuttings or grafting has been a key element in the domestication of long-lived perennials, allowing the maintenance and spread of superior individuals despite self-incompatibility . However, the use of such techniques has further decreased the number of sexual cycles in tree crops since the initial domestication event, adding to the effect of long juvenile phases in limiting the genetic divergence between cultivated trees and their wild progenitors , , , . Thus, domestication can generally be considered more recent, at least in terms of the number of generations, in fruit tree crops than in seed-propagated selfing annuals.
Given the slow process of selection and the limited number of generations in which humans could exert selection, the protracted nature of the domestication process in trees has probably resulted in limited bottlenecks ,
and in a weaker domestication syndrome
than in seed-propagated annuals. Nevertheless, many cultivated fruit trees clearly display morphological, phenotypic and physiological features typical of a domestication syndrome, such as large fruits and high sugar or oil content , . Many aspects of fruit tree domestication have been little studied . Consequently, most of the hypotheses concerning the consequences of particular features of trees for their domestication/diversification remain to be tested. Recent studies on grapevines, almond and olive trees have provided illuminating insights, such as the importance of outcrossing and interspecific hybridization , , , but additional studies of other species are required to draw more general conclusions.
Here, we investigated the origins of the domesticated apple Malus domestica Borkh., one of the most emblematic and widespread fruit crops in temperate regions . A form of apple corresponding to extant domestic apples appeared in the Near East around 4,000 years ago , at a time corresponding to the first recorded uses of grafting. The domesticated apple was then introduced into Europe and North Africa by the Greeks and Romans and subsequently spread worldwide . While the ancestral progenitor has been clearly identified as being M. sieversii, the identity and relative contributions of other wild species present along the Silk route that have contributed to the genetic makeup of apple cultivars remain largely unknown. This is surprising given the potential importance of this knowledge for plant breeding and for our understanding of the process of domestication in fruit trees.
The wild Central Asian species M. sieversii (Ldb.) M. Roem has been identified as the main contributor to the M. domestica genepool based on similarities in fruit and tree morphology, and genetic data , , , . The Tian Shan forests were identified as the geographic area in which the apple was first domesticated, on the basis of the considerable intraspecific morphological variability of wild apple populations in this region , . Nucleotide variation for 23 DNA fragments even suggested that M. sieversii and M. domestica belonged to a single genepool (which would be called M. pumila Mill.), with phylogenetic networks showing an intermingling of individuals from the two taxa . Some authors have also suggested possible contributions of additional wild species present along the Silk Route: M. baccata (L.) Borkh, which is native to Siberia, M. orientalis Uglitz., a Caucasian species present along western sections of the ancient trade routes, and M. sylvestris Mill. (European crabapple), a species native to Europe , , , . These hypotheses were based on the history of human migration and trade, the lack of phylogenetic resolution between M. domestica and these four wild species , , genetic evidence of hybridization at a local scale between domesticated apple and M. sylvestris , and the recent finding of sequence haplotype sharing between M. sylvestris and M. domestica . However, such secondary contributions remain a matter of debate, mostly due to the difficulty of distinguishing introgression from incomplete lineage sorting , , . The three wild species occurring along the Silk Route all bear small, astringent, tart fruits. None of these species has the fruit quality of M. sieversii, but they may have contributed other valuable horticultural traits, such as later flowering, resistance to pests and diseases, capacity for longer storage or climate adaptation. The organoleptic properties of the fruits of these wild species may also have been selected during domestication, for the preparation of apple-based beverages, such as ciders , . Cider apples are indeed smaller, bitter and more astringent than dessert apples and bear some similarity to M. sylvestris apples. There is also evidence to suggest that Neolithic and Bronze Age Europeans were already making use of M. sylvestris .
In this study, we used a comprehensive set of apple accessions sampled across Eurasia (839 accessions from China to S ) and 26 microsatellite markers distributed evenly across the genome to investigate the following questions: 1) Is there evidence for population subdivision within and between the five taxa M. domestica, M. baccata, M. orientalis, M. sieversii and M. sylvestris? 2) How large is the contribution of wild species other than the main progenitor, M. sieversii, to the genome of M. domestica? 3) Does M. domestica have a genetic structure associated with its different possible uses (i.e., differences between cider and dessert apples)? 4) What consequences have domestication, subsequent crop improvement and vegetative propagation by grafting had for genetic variation in cultivated apples? Most of our samples of M. domestica corresponded to cultivars from Western Europe ( and ), as almost all the cultivars available in modern collections (including American, Australasian cultivars) are of European ancestry and this region is therefore the most relevant area for the detection of possible secondary introgression from the European crabapple.
Geographic origins of the samples of the four wild Malus species used: M. sylvestris (blue), M. orientalis (yellow), M. baccata (purple), and M. sieversii (red).Samples of unknown origin (N = 28) were not projected onto the map.
High diversity and low deviations from random mating expectations within species
Our sampling scheme ( and ), based on the collection of a single tree for each apple variety, was designed to avoid the sampling of clones. However, there may still be some clonality if some varieties differing by only a few mutations were propagated by grafting. We corrected for this potential clonality, using the clonal assignment procedures implemented in GENODIVE . We found no pair of samples assigned to the same clonal lineage unless using a threshold of 22 pairwise differences between multilocus genotypes, indicating that our samples did not include any clonal genotypes (the threshold corresponds to the maximum genetic distance allowed between genotypes deemed to belong to the same clonal lineage).
Many apple cultivars, including modern cultivars in particular, share recent common ancestors, and siblings or clones of wild species can also be collected unintentionally in the field. Because these features could result in a spurious genetic structure due to the presence of closely related individuals in the dataset, we checked for the presence of groups of related individuals in our dataset between M. domestica cultivars and between the individuals of each wild species. The percentage of pairs with a pairwise relatedness (rxy) greater than 0.5 (i.e., full sibs) was: 0.4% in M. domestica (N = 168 pairs), 0.3% in M. sieversii (N = 79), 0.004% in M. orientalis (N = 20), and 0.7% in M. baccata (N = 40). For M. sylvestris, no individual pair with rxy&0.5 was identified. However, the distribution of pairwise relatedness rxy among M. domestica cultivars did not deviate significantly from a Gaussian distribution centred on 0 and with a low variance (Fisher's exact test, P≈1, standard deviation = 0.11, ). This suggests that closely related cultivars are unlikely to have biased subsequent analyses of population structure. We also checked for the limited effect of relatedness on our conclusions by performing all analyses of population subdivision on both the full dataset and a pruned dataset excluding related individuals (see below).
We tested the null hypothesis of random mating within each species by calculating FIS, which measures inbreeding. All five Malus species had relatively low values of FIS, although all were significantly different from zero (), suggesting that each species corresponded to an almost random mating unit. This is consistent with the self-incompatibility system of these species and indicates a lack of widespread groups of related individuals in M. domestica. Low FIS values at species level also indicate a lack of population structure within species. The higher values of FIS observed in M. baccata probably resulted from the occurrence of null alleles, as the microsatellite markers were developed in M. domestica, to which M. baccata is the most distantly related (). The lowest FIS value was that obtained for M. domestica, reflecting outcrossing between dissimilar parents in breeding programs, or that selection targeted higher levels of heterozygosity .
Summary of genetic variation in the five Malus species.
Pairwise differentiation (FST) between the five Malus species.
The five Malus species form well separated genetic clusters
We used the ‘admixture model’ implemented in STRUCTURE 2.3
to infer population structure and introgression. Analyses were run for population structure models assuming K = 1 to K = 8 distinct clusters (). The ΔK statistic, designed to identify the most relevant number of clusters by determining the number of clusters beyond which there is no further increase in likelihood , was greatest for K = 3 (ΔK = 6249, Pr|ln L = -78590). However, the clusters identified at higher K values may also reveal a genuine and biologically relevant genetic structure, provided that they are well delimited . The five Malus species were clearly assigned to different clusters for models assuming K≥6 clusters and for a minor clustering solution (“mode”) at K = 5 (). The major mode (i.e., the clustering solution found in more than 60% of the simulation replicates) observed at K = 5 grouped together M. sylvestris and M. domestica genotypes. Increasing the number of clusters above K = 6 identified no additional well-delimited clusters corresponding to a subdivision of a previous cluster. Instead, it simply introduced heterogeneity into membership coefficients, indicating that the clustering of the five Malus species into separate genepools was the most relevant clustering solution. We checked that the presence of related pairs of cultivars in our dataset did not bias clustering results, by repeating the analysis on a pruned dataset (N = 489) excluding all related individuals in wild and cultivated species (i.e., excluding all pairs with rxy≥0.5). Similar results were obtained, with the same five distinct clusters identified as for the full dataset.
Proportions of ancestry of Malus genotypes from five species (N = 770) from K = 2 to K = 8 ancestral genepools (“clusters”) inferred with the STRUCTURE program.Each individual is represented by a vertical bar, partitioned into K segments representing the amount of ancestry of its genome in K clusters. When several clustering solutions (“modes”) were represented within replicate runs, the proportion of simulations represented by each mode is given.
We estimated the genetic differentiation between the five Malus species by calculating pairwise FST (). All FST values were highly significant (P&0.001) and seemed to indicate a West to East differentiation gradient of M. domestica with the wild species. The highest level of differentiation was that between M. baccata and the other Malus species, and the lowest level of differentiation was that between M. domestica and the westernmost species, M. sylvestris (). Malus domestica was markedly more differentiated from its main progenitor M. sieversii (FST = 0.0639) than from the European M. sylvestris (FST = 0.006) and it was only slightly less differentiated from the Caucasian M. orientalis (FST = 0.049).
No bottleneck during apple domestication
We first searched for footprints of a domestication bottleneck by comparing levels of microsatellite variation in M. domestica and wild species. There was no significant difference in genetic diversity (as measured by expected heterozygosity, HE) between M. domestica and M. baccata, M. orientalis or M. sieversii, but HE was significantly higher in M. sylvestris than in M. domestica (). Significant differences in allelic richness (Ar) were found between M. domestica and M. orientalis (Wilcoxon signed rank test, P = 0.03) or M. sylvestris (P&10-8), but not between M. domestica and either M. baccata (P = 0.9) or M. sieversii (P = 0.9) ().
We used the method implemented in the BOTTLENECK program , comparing the expected heterozygosity estimated from allele frequencies with that estimated from the number of alleles and the sample size, which should be identical for a neutral locus in a population at mutation-drift equilibrium. Inferences about historical changes in population size are based on the prediction that the expected heterozygosity estimated from allele frequencies decreases faster than that estimated under a given mutation model at mutation-drift equilibrium in populations that have experienced a recent reduction in size. BOTTLENECK analysis showed no significant deviation from mutation-drift equilibrium in any of the five species, under either stepwise or two-phase models of microsatellite evolution (one-tailed Wilcoxon signed rank test, P&0.95). We therefore detected no signal of a demographic bottleneck associated with the domestication of apples.
Variable recent contributions of wild relative species to the M. domestica genepool, with the strongest introgression from M. sylvestris
We used the admixture coefficients estimated by STRUCTURE to assess the recent contribution of the various wild species to the M. domestica genepool. STRUCTURE analyses of the full dataset showed some admixture among Malus species for the minor mode separating the five species at K = 5. Admixture coefficients were higher between M. domestica and M. sylvestris (α = 0.23) than between M. domestica and respectively M. sieversii (α = 0.06), M. orientalis (α = 0.034) and M. baccata (α = 0.032).
We further analysed the contribution of each wild species to the genome of M. domestica by running STRUCTURE separately on each pair of species including M. domestica (;
and ). Malus domestica genotypes with membership coefficients ≥0.20 in a wild species genepool were considered to display introgression. Using this somehow arbitrary cut-off value, STRUCTURE analyses revealed that 26% of M. domestica cultivars displayed introgression from the European crabapple, M. sylvestris ( and ). By contrast, only 2%, 3% and 0.02% of the M. domestica genotypes displayed introgression from M. sieversii, M. orientalis and M. baccata, respectively ( and ). The M. domestica cultivars displaying admixture with the M. sylvestris genepool were mostly Russian (e.g., “Antonovka”, “Antonovka kamenicka”, “Novosibirski Sweet”, “Yellow transparent”), French (e.g., “Blanche de St Anne”, “St Jean”, “Api” and “Michelin”) and English (e.g., “Worcester Pearmain” and “Fiesta”). The M9 dwarf apple cultivar (“Paradis jaune de Metz”, ) commonly used as a rootstock also appeared to display introgression from the European crabapple (proportion of ancestry in the M. domestica genepool: 0.28; ). When French cultivars were removed from the dataset (N = 89) and pairwise STRUCTURE analyses were repeated for all species pairs including M. domestica, 18% of cultivars displayed introgression from M. sylvestris, including commercial cultivars such as Granny Smith, Michelin, Antonovka and Ajmi () with a mean membership coefficient of M. sylvestris into M. domestica genepool of 47%. Malus sylvestris thus appears to have made a significant contribution to the M. domestica genepool through recent introgression, building on the more ancient contribution (see below) of the Asian wild species M. sieversii. We also note that a few M. domestica individuals appeared to display introgression from several wild species (), and that M. baccata ornamental cultivars, such as M. baccata flexilis, M. baccata Hansen's and M. baccata gracilis, were partially or even mostly assigned (from 32% to &80%) to the M. domestica genepool ().
Proportions of ancestry in two ancestral genepools inferred with the STRUCTURE program, based on datasets including M. domestica (green, N = 299) and each of the four wild Malus species (red).The x-axis is not to scale (details in ).
Mean proportions of assignment to each of the two species in species pair comparisons (K = 2) including M. domestica (Genepool 1) and each of the four wild Malus species (Genepool 2).
Wild Central Asian apple origin of the M. domestica genepool
Previous studies , ,
identified the Central Asian wild apple M. sieversii as the main progenitor of M. domestica on the basis of DNA sequences. Due to the large contribution by M. sylvestris detected in our dataset, corresponding mostly to Western European cultivars, M. domestica and M. sylvestris appeared to be the most closely related pair of species in our analyses of microsatellite markers. We investigated the more ancient contribution of M. sieversii to the M. domestica genepool, by reassessing the genetic differentiation between species in analyses restricted to “pure” individuals (i.e., assigned at ≥0.9 to their respective genepools) from both wild and cultivated species. All FST values were highly significant (P&0.001), but the ranking of FST values between M. domestica and the various wild species was affected: the highest differentiation was still observed between M. domestica and M. baccata (FST = 0.22), but the lowest differentiation was observed between M. domestica and M. sieversii (FST = 0.11). Regarding the differentiation between M. sylvestris and M. domestica, we observed the opposite of what was found with the full dataset: M. sylvestris appeared to be more strongly differentiated (FST = 0.14) from M. domestica than M. sieversii. Thus, by removing signals of recent introgression between cultivated and wild species we were able to confirm that M. sieversii was the initial progenitor of M. domestica.
Recent introgression from M. domestica into wild species
The finding of a significant level of introgression from wild species into cultivated apple suggested that gene flow might also have occurred in the opposite direction. STRUCTURE analyses of pairs of species confirmed this hypothesis (), revealing possible introgression of genetic material into M. sylvestris, M. baccata, M. orientalis and M. sieversii from M. domestica (mean proportions of ancestry in the M. domestica genepool of 0.12, 0.10, 0.03 and 0.23, ). Considering genotypes with membership coefficients ≥0.9 in the M. domestica genepool as misclassified, we found a total of N = 31 misclassified wild Malus individuals. These results suggest gene flow from the domesticated apple genepool could significantly affect the genetic integrity of wild apple relatives, their future evolution and, possibly, their use as resources for crop improvement.
Inference of demographic history
Model-based Bayesian clustering algorithms, such as that implemented in STRUCTURE, have a high level of power only for the detection of recent introgression events , , . We therefore investigated the contributions of M. sylvestris and M. orientalis to the M. domestica genepool using approximate Bayesian computation (ABC) methods that offer a more historical perspective on gene flow . We used a demographic model implementing admixture events .
We compared several admixture models to infer what species pairs underwent introgression events and to estimate introgression rates . Malus baccata was not included in these analyses because of its high level of divergence from M. domestica. We assumed, as suggested by previous studies, that M. domestica derived originally from M. sieversii. The most complex model simulated sequential admixtures between M. domestica and all wild species. Other models sequentially removed introgression with each wild species, the order being based on FST values and admixture rates inferred by STRUCTURE. The compared models were the following: (i) the model a assumed that M. domestica was derived from M. sieversii and that the ancestral M. domestica population was involved in reciprocal introgression events with M. orientalis and M. sylvestris, and subsequently introgressed back into M. sieversii (), (ii) model b was similar to the model a, but without introgression events from M. domestica into wild species (), (iii) the model c included a single introgression event, from M. sylvestris into M. domestica (), and (iv) the model d simulated no admixture (). The number of parameters estimated in the model was limited by fixing the times of admixture with M. orientalis, M. sylvestris and M. sieversii at 600, 200 and 13 generations before the present, respectively. We used the following underlying hypotheses: (i) as the juvenile period of Malus lasts five to 10 years, we assumed a generation time of 7.5 years, (ii) admixture between ancestral M. domestica and M. orientalis in the Caucasus occurred approximately 4,500 years ago, shortly before the appearance of sweet apples in the Middle East (4,000 years ago), (iii) admixture between ancestral M. domestica and M. sylvestris in Europe occurred approximately 1,500 years ago, soon after the introduction of domesticated apples into Europe by the Greeks and Romans (iv) back-introgression into M. sieversii from M. domestica occurred approximately 100 years ago, when the cultivation of modern varieties reached Central Asia.
Admixture models compared in approximate Bayesian computations.Model a assumes that M. domestica is derived from M. sieversii and that the ancestral M. domestica population was involved in reciprocal introgression events with M. orientalis and M. sylvestris, and subsequently introgressed back into M. sieversii. Model b assumes no introgression from M. domestica into wild species, model c assumes the only admixture event is from M. sylvestris into M. domestica, and model d assumes no admixture. Admixture times between M. domestica and the three wild species were fixed (see text). Abbreviations: Nk, effect Tk, r1, r3, r4 introgression from M. domestica into M. sieversii, M. sylvestris, and M. orientalis r2, r5 introgression from M. sylvestris and M. orientalis, respectively, into M. domestica.
The relative posterior probabilities computed for each model provided strongest statistical support for model c, which assumed a single introgression event, from M. sylvestris into M. domestica (; posterior probability [p] = 0.67, 95% confidence interval: 0.63–0.72). Note that the model without admixture (model d) had the lowest relative posterior probability (). In analyses under alternative admixture models (models a and b), the posterior distributions were flat for introgression between M. domestica and M. orientalis and highly skewed towards low values for introgression into M. sylvestris and M. sieversii (not shown), which is consistent with statistical support being highest for model c.
Relative posterior probabilities (p) for the four historical models compared using approximate Bayesian computations.Given that the model c was clearly favoured, parameter estimates are shown below only for this model (; prior distributions in ). The contribution of M. sylvestris to the M. domestica genepool was estimated at about 61% (95% credibility interval [95% CI]: 50–68%). We obtained estimates of effective population sizes of 3,520 (95% CI: 2,090–5,680) for M. domestica, 13,200 (95% CI: 6,920–19,300) for M. sieversii, 34,600 (95% CI: 15,100–48,000) for M. sylvestris, and 28,300 (95% CI: 11,700–64,000) for M. orientalis. Using a generation time of 7.5 years, the divergence between M. domestica and M. sieversii (T3) was estimated to have occurred 17,700 years ago (95% CI: 6,225–25,200), which is earlier than previously thought, but we note that the credibility interval is quite large. We estimated that M. sylvestris and M. sieversii diverged about 83,250 years ago (T1, 95% CI: 40,575–334,500), with M. orientalis and M. sieversii diverging about 20,775 years ago (T2, 95% CI: 9,900–47,775).
Demographic and mutation parameters estimated using approximate Bayesian computation for model c.The results above were obtained using the full dataset. We checked the validity of our inferences by conducting analyses on the dataset without admixed and misclassified individuals and using different times of admixture, by assessing the goodness-of-fit of models to data, and by checking that sufficient power was achieved to discriminate among competing models (; , , ). Overall, ABC analyses all provided clear support for a model with contribution of the European crabapple into the domesticates, although the estimated value of the actual contribution of M. sylvestris is probably overestimated here, and should therefore be treated with caution. Indeed, the simulation of a single introgression event hundreds of years ago most likely demanded higher rates of introgression to account for the actual genetic contribution of M. sylvestris into M. domestica than would be needed under continuous gene flow over a long period.
Weak genetic structure within M. domestica: linked to cultivar use or geography?
As cider cultivars produce apples that are smaller, more bitter and astringent than dessert cultivars, we expected to observe genetic differentiation between these two groups of cultivars and a closer genetic proximity of cider cultivars to M. sylvestris , . Neither hypothesis was supported by our data. The classification of apples into “dessert” and “cider” varieties as prior information for STRUCTURE (Locprior model) revealed a very weak tendency of cider and dessert cultivars to be assigned to different clusters at K = 2 (), but increasing K did not further result in clearer differentiation between the two types of cultivars. At K = 2, M. domestica cider genotypes had a mean membership of 94.7%, and M. domestica dessert genotypes had a mean membership of 52.5%. However, STRUCTURE analyses without this prior information gave essentially the same clustering patterns at K = 2 (G′ = 0.95 similarity to analyses using classification to assist clustering). The weak differentiation between cider and dessert cultivars (FST = 0.02) and their high level of admixture in STRUCTURE analyses () indicated a shallow subdivision of the M. domestica genepool. Analyses on a pruned dataset from which closely related individuals had been removed (i.e., pairs of genotypes with rxy≥0.5; N = 172) revealed the same pattern, confirming that the presence of related cultivars in the dataset did not bias clustering analyses. STRUCTURE was also run on a dataset including all M. sylvestris genotypes, to test the hypothesis that cider cultivars would display a higher level of introgression from the European crabapple. However, the opposite pattern was observed: the proportion of genotypes displaying introgression from M. sylvestris was actually significantly higher in dessert than in cider cultivars (36.4% and 15.5% respectively, χ2 = 16.9, P = 4×10-5). Finally, little genetic differentiation was observed between groups of cultivars of different geographic origins (95% CI: -0.8–0.6, ).
Proportions of ancestry of M. domestica genotypes (cider and dessert apples) in two ancestral genepools inferred with the STRUCTURE program.
DiscussionThe apple is so deeply rooted in the culture of human populations from temperate regions that it is often not recognized as an exotic plant of unclear origin. We show here that the evolution of the domesticated apple involved more than one geographically restricted wild species. The domesticated apple did not arise from a single event over a short period of time, but from evolution extending over thousands of years. The genepool of the current domesticated apple varieties has been enriched by the contribution of at least two wild species. Malus species have a self-in apple domestication and traditional variety improvement have therefore been based mostly on the selection of the best phenotypes grown from open-pollinated seeds. This breeding strategy has probably favoured the incorporation of genetic material from multiple wild sources and the maintenance of high levels of genetic variation in domesticated apples, despite the extensive use of large-scale vegetative propagation of superior individuals by grafting. Our results are consistent with those reported for the few other woody perennials studied to date, such as grape , red mombin
and olive trees , and support the view that domestication in long-lived plants differs in many respects from the scenarios described for seed-propagated annuals.
Weak differentiation from wild progenitors and the Central Asian origin of M. domestica
Malus sieversii was previously identified as the main contributor to the M. domestica genome on the basis of morphological and sequence data , . The flanks of the Tian Shan mountains have been identified as a likely initial site of domestication, based on the high morphological variability of the wild apples growing in this region, and their similarity to sweet dessert apples , . We show here, using a set of rapidly evolving genetic markers distributed throughout the genome and a large sampling, that M. domestica now forms a distinct, random mating group, surprisingly well separated from M. sieversii, with no difference in levels of genetic variation between the domesticate and its wild progenitor. This contrasts with the pattern previously reported, based on a twenty three-gene phylogenetic network , where domesticated varieties of apple appeared nested within M. sieversii. After the removal of individuals showing signs of recent admixture, M. sieversii and M. domestica nevertheless appeared to be the pair of species most closely related genetically, confirming their progenitor-descendant relationship.
Lack of a domestication bottleneck
Apple breeding methods (grafting and “chance seedling” selection), life-history traits specific to trees and/or the genetic architecture of selected traits have likely played a role in the conservation of levels of genetic diversity in cultivated apples similar to those in wild apples. Some factors, such as “chance seedling” selection , may even have increased genetic diversity, by favouring outcrossing events among domesticates and introgression from wild species . The low inbreeding coefficients inferred in domesticated apples and the low level of differentiation between cultivated and wild apple populations , , ,
indicate a high frequency of crosses between individuals of M. domestica, M. sieversii and other wild relatives hailing from diverse geographic origins. Such a high level of gene flow has likely contributed to maintenance of a high level of genetic diversity in domesticated apples.
The grafting technique, which was probably developed around 3,000 years ago, has made it possible to propagate superior individuals clonally. The spread of grafting, together with the lengthy juvenile phase (5–10 years) and the long lifespan of apples, may have imposed strong limits on the intensity of the domestication bottleneck thereby limiting the loss of genetic diversity , , . By decreasing the number of generations since domestication, these factors have probably also helped to restrict the differentiation between domesticates and wild relatives. In theory, grafting may have limited the size of the apple germplasm dispersed early on to a few very popular genotypes, thereby provoking a sudden shrink in effective population size and a loss of diversity. However, we found no evidence that the clonal propagation of apples resulted in a long-lasting decrease in population size or clonal population structure. We can speculate that this may be due to a combination of various factors such as: gene flow with wild species, small-scale propagation (many farmers producing a few grafts each), a large variation in preferences for taste and other quality characteristics between farmers and cultures, large differences in growth conditions leading to the adoption of different sets of genotypes in different regions or the typical behaviour of hobby breeders, who tend to spot particular differences and multiply them. Similarly, for grape, there are huge numbers of old varieties and as much genetic variation in cultivated varieties as in wild-relative progenitors .
A major secondary contribution from the European crabapple
There has been a long-running debate concerning the possible contribution of other wild species present along the Silk Route to the genetic makeup of M. domestica , , , , . Our results clearly show that interspecific hybridization has been a potent force in the evolution of domesticated apple varieties. Apple thus provides a rare example of the evolution of a domesticated crop over a long period of time and involving at least two wild species (see also the cases of olive tree and avocado , , , ). A recent study argued that introgression from M. sylvestris into the M. domestica genepool was the most parsimonious explanations for shared gene sequence polymorphisms between the two species . Using an unprecedentedly large dataset, more numerous and more rapidly evolving markers and a combination of inferential methods, we provide a comprehensive view of the history of domestication in apple. We confirm that M. sieversii was the initial progenitor and show that the wild European crabapple M. sylvestris has been a major secondary contributor to the diversity of apples, resulting in current varieties of M. domestica being more closely related to M. sylvestris than to their central Asian progenitor. This situation is reminiscent of that for maize, in which the cultivated crop Zea mays is genetically more closely related to current-day highland landraces than to lowland Z. mays ssp. parviglumis from which the crop was domesticated . This pattern has been attributed to large-scale gene flow from a secondary source, a second subspecies of teosinte, Z. mays ssp. mexicana, into highland maize populations .
The usefulness of wild relatives for improving elite cultivated crop genepools has long been recognised and the exploitation of wild resources is now considered a strategic priority in breeding and conservation programs for most crops , , . Domesticated apples are unusual in that the contribution of wild relatives probably occurred early and unintentionally in the domestication process, preceding even the use of controlled crosses. The use of genetic markers with lower mutation rates than our set of microsatellites might also make it possible to investigate the contribution of more phylogenetically distant apple species growing in areas away from the Silk Route to the diversification of modern apple cultivars.
The Romans introduced sweet apples into Europe at a time at which the Europeans were undoubtedly already making cider from the tannin-rich fruits of the native M. sylvestris , . Cider is not typical of Asia , but it was widespread in Europe by the time of Charlemagne (9th century, ). Large numbers of apple trees were planted for cider production in France and Spain from the 10th century onwards , . The very high degree of stringency of cider apples (often to the extent that they are inedible) led to the suggestion that cider cultivars arose from hybridization between M. sylvestris and sweet apples , , . We show here that the genetic structure within the cultivated apple genepool is very weak, with poor differentiation between cider and dessert apples. Cider cultivars thus appear to be no more closely genetically related to M. sylvestris than dessert cultivars. As wild Asian apples are known to cover the full range of tastes , , it is possible that fruits with the specific characteristics required for cider production were in fact initially selected in Central Asia and subsequently brought into Europe. There is a long-standing tradition of cider production in some parts of Turkey , for instance, which is potentially consistent with an Eastern origin of cider cultivars. However, the low level of genetic differentiation between dessert and cider apples indicates that, even if different types of apples were domesticated in Asia and brought to Europe, they have not diverged into independent genepools.
Concluding remarks
This study settles a long-running debate by confirming that 1) M. domestica was initially domesticated from M. sieversii, and 2) M. domestica subsequently received a significant genetic contribution from M. sylvestris, much larger than previously suspected , at least in Western Europe, where originated most of our samples and most cultivar diversity. The higher level of introgression of the European crabapple into the domesticated apple in this study than in previous studies , ,
may be attributed to the use of a larger and more representative set of M. domestica genotypes coupled with the genotyping of numerous and rapidly evolving markers known to trace back more recent events.
Our inferences also have important implications for breeding programs and for the conservation of wild species of apple. The major contribution of the various wild species to the M. domestica genepool highlights the need to invest efforts into the conservation of these species, which may contain unused genetic resources that could further improve the domesticated apple germplasm , such as disease resistance genes or genes encoding specific organoleptic features.
Materials and Methods
Sample collection and DNA extraction
Leaf material was retrieved from the collections of various institutes (INRA Angers, F USDA - ARS, Plant Genetic Resources Unit, Geneva, NY; ILVO Melle, Belgium) and from a private apple germplasm repository in Brittany for M. domestica (N = 368,
including only diploid cultivars N = 299) and from forests for the four wild species (; ). Malus sieversii (N = 168) material was collected from 2007 to 2010 in the Chinese Xinjiang province (N = 26), Kyrgyzstan (N = 5), Uzbekistan (N = 1), Tajikistan (N = 1) and Kazakhstan (N = 114). Malus orientalis (N = 215) was sampled in 2009 in Armenia (N = 203), Turkey (N = 5) and Russia (N = 5). Malus sylvestris (N = 40) samples were obtained from 15 European countries. Malus baccata (N = 48) was sampled in 2010 in Russia. The origins of M. domestica cultivars were: France (N = 266), Great Britain (N = 12), USA (N = 12), Russia (N = 7), the Netherlands (N = 6), Australia (N = 4), Belgium (N = 4), Germany (N = 4), Japan (N = 3), Ukraine (N = 3), Tunisia (N = 2), Switzerland (N = 2), Spain (N = 2), New Zealand (N = 2), Israel (N = 1), Ireland (N = 1), Canada (N = 1), Armenia (N = 2) and unknown/debated (N = 34). Genomic DNA was extracted with the Nucleo Spin plant DNA extraction kit II (Macherey & Nagel, Düren, Germany) according to the manufacturer's instructions.
Microsatellite markers and polymerase chain reaction (PCR) amplification
Microsatellites were amplified by multiplex PCR, with the Multiplex PCR Kit (QIAGEN, Inc.). We used 26 microsatellites spread across the 17 chromosomes (one to three microsatellites per chromosome), in 10 different multiplexes previously optimised on a large set of genetically related progenies of M. domestica . The four multiplexes (MP01, MP02, MP03, MP04; ; Lasserre P. unpublished data) were performed in a final reaction volume of 15 ul (7.5 ul of QIAGEN Multiplex Master Mix, 10–20 uM of each primer, with the forward primer labelled with a fluorescent dye and 10 ng of template DNA). We used a touch-down PCR program (initial annealing temperature of 60°C, decreasing by 1°C per cycle down to 55°C). Six other multiplex reactions (Hi6, Hi4ab, Hi5-10, Hi13a, Hi13b, Hi4b) were performed using previously described protocols . Genotyping was performed on an ABI PRISM X3730XL, with 2 ul of GS500LIZ size standard (Applied Biosystems). Alleles were scored with GENEMAPPER 4.0 software (Applied Biosystems). We retained only multilocus genotypes presenting less than 30% missing data.
Suitability of microsatellites for population genetic analyses
We checked the suitability of the markers for population genetic analyses. None of the 26 microsatellite markers deviated significantly from a neutral equilibrium model, as shown by the non significant P-values obtained in Ewen-Watterson tests , and no pair of markers was found to be in significant linkage disequilibrium in any of the species , . The markers could therefore be considered unlinked and neutral.
Analyses of genetic variation and differentiation between the five species
Apple cultivars may be polyploid . We therefore first checked for the presence of polyploidy individuals of M. domestica within our dataset. Individuals presenting multiple peaks on electrophoregrams were first re-extracted to eliminate contamination as a possible source of apparent polyploidy. We then checked whether they had been reported to be polyploidy in previous studies . After completion of this checking procedure, we removed 69 polyploids (of the 368 samples) from subsequent analyses. We tested for the occurrence of null alleles at each locus with MICROCHECKER 2.2.3 software . Allelic richness and private allele frequencies were calculated with ADZE software , for a sample size of 22. Heterozygosity (expected (HE) and observed (HO)), Weir & Cockerham F-statistics, deviation from Hardy-Weinberg equilibrium and genotypic linkage disequilibrium were estimated with GENEPOP 4.0 , . The significance of differences between FST values was assessed in exact tests carried out with GENEPOP 4.0 , . Individuals were assigned to clonal lineages with GENODIVE . We estimated relatedness between pairs of cultivars and between pairs of individuals within each species, by calculating the rxy of Ritland and Lynch
with RE-RAT online software . We tested whether the distributions of rxy deviated significantly from a Gaussian distribution with a mean of zero and a standard deviation equal to the observed standard deviation, by comparing observed and simulated distributions in Fisher's exact test (R Development Core Team, URL ).
Assessing bottlenecks during apple domestication and diversification
We tested for the occurrence of a bottleneck during apple domestication with the method implemented in BOTTLENECK , . The tests were performed under the stepwise-mutation model (SMM) and under a two-phase model (TPM) allowing for 30% multistep changes. We used Wilcoxon signed rank tests to determine whether a population had a significant number of loci with excess genetic diversity.
Analyses of population subdivision
We used the individual-based Bayesian clustering method implemented in STRUCTURE 2.3.3 , ,
to investigate species delimitation, intraspecific population structure and admixture. This method is based on Markov Chain Monte Carlo (MCMC) simulations and is used to infer the proportion of ancestry of genotypes in K distinct predefined clusters. The algorithm attempts to minimize deviations from Hardy–Weinberg and linkage equilibrium within clusters. Analyses were carried out without the use of prior information, except for analyses of population subdivision within the M. domestica genepool for which the “cider”/“dessert” classification of cultivars was used as prior information to assist clustering. K ranged from 1 to 8 for analyses of the five-species dataset and the M. domestica dataset, and was fixed at K = 2 for analyses of pairs of species including M. domestica and each of the wild species. Ten independent runs were carried out for each K and we used 500,000 MCMC iterations after a burn-in of 50,000 steps. We used CLUMPP v1.1.2 (Greedy algorithm)
to look for distinct modes among the 10 replicated runs of each K.
STRUCTURE analyses were run for the full dataset (N = 839) and for two pruned datasets excluding non-pure individuals (i.e., genotypes with &0.9 membership of their species' genepool) and related individuals (rxy≥0.5).
Inference of demographic history
We used the DIYABC program
to compare different admixture models and infer historical parameters. We simulated microsatellite datasets for 14 loci (Ch01h01, Ch01h10, Ch02c06, Ch02d08, Ch05f06, Ch01f02, Hi02c07, Ch02c09, Ch03d07, Ch04c07, Ch02b03b, MS06g03, Ch04e03, Ch02g01) previously reported to be of the perfect repeat type , , . In total, we generated 5×105 simulated datasets for each model.
A generalized stepwise model (GSM) was used as the mutational model. The model had two parameters: the mean mutation rate (μ) and the mean parameter (P) of the geometric distribution used to model the length of mutation events (in numbers of repeats). As no experimental estimate of microsatellite mutation rate is available for Malus, the mean mutation rate was drawn from a uniform distribution by extreme values of 10-4 and 10-3, and the mutation rate of each locus was drawn independently from a Gamma distribution (mean = μ; shape = 2). The parameter P ranged from 0.1 to 0.3. Each locus L had a possible range of 40 contiguous allelic states (44 for CH02C06, 42 for CH04E03) and was characterized by individual values for mutation rate (μL) and the parameter of the geometric distribution (PL); μL and PL were drawn from Gamma distributions with the following parameter sets: mean = μ, shape = 2, range = 5×10-5–5×10-2 for μL, and mean = P, shape = 2, range = 0.01–0.9 for PL. As not all allele lengths were multiples of motif length, we also included single-nucleotide insertion-deletion mutations in the model, with a mean mutation rate (μSNI) and locus-specific rates drawn from a Gamma distribution (mean = μSNI; shape = 2). The summary statistics used were: mean number of alleles per locus, mean genetic diversity , genetic differentiation between pairwise groups (FST; ), genetic distances (δμ)2 .
We used a polychotomous logistic regression procedure
to estimate the relative posterior probability of each model, based on the 1% of simulated data sets closest to the observed data. Confidence intervals for the posterior probabilities were computed using the limiting distribution of the maximum likelihood estimators . Once the most likely model was identified, we used a local linear regression to estimate the posterior distributions of parameters under this model . The 1% simulated datasets most closely resembling the observed data were used for the regression, after the application of a logit transformation to parameter values.
Supporting InformationGeographic origins of diploid M. domestica cultivars (N = 299). See details in .
Distribution of pairwise relatedness coefficients
among the M. domestica cultivars. rxy values among cultivars are normally distributed around a mean of zero, with a low variance between pairs of cultivars (Fisher's exact test, P≈1).
Proportions of ancestry in two ancestral genepools inferred with the STRUCTURE program from datasets including M. domestica (green, N = 89) and each of the wild Malus species (red) except M. baccata. The x-axis is not at scale.
Description of the Malus species accessions analysed, with their geographic origin and providers.
Malus domestica cultivars used in the study, with their use, provider, geographic putative origin. Details of the STRUCTURE analysis summarized in
are also provided.
Membership coefficients inferred from the STRUCTURE analysis for M. baccata individuals.
Prior distributions used in approximate Bayesian computations. Prior distributions are uniform between lower and upper bound. Parameters are introduced in
and . Species names are abbreviated.
Relative posterior probabilities (p) for the four historical models compared using approximate Bayesian computations. Models are described in . CI2.5 and CI97.5 are boundaries of the 95% confidence intervals. (A) Analyses on a pruned dataset with misclassified wild individuals and individuals with a recent admixed ancestry removed. (B) Analyses on the full dataset, assuming that admixture between ancestral M. domestica and M. sylvestris was more recent (67 generations–500 ybp) than in original analyses (200 generations–1,500 ybp).
Demographic and mutation parameters estimated using approximate Bayesian computation for model c. Posterior distributions are summarized as the mode and boundaries of the 95% credibility intervals (CI2.5 and CI97.5). Demographic parameters are introduced in
(note that admixture times are fixed in these analyses). Composite parameters scaled by the mutation rate are also shown. The mutation parameters are μ (mean mutation rate), p (mean value of the geometric distribution parameter that governs the number of repeated motifs that increase or decrease the length of the locus during mutation events), μSNI (mean single nucleotide indel mutation rate). Species names are abbreviated. (A) Analyses on a pruned dataset with misclassified wild individuals and individuals with a recent admixed ancestry removed. (B) Analyses on the full dataset, assuming that admixture between ancestral M. domestica and M. sylvestris was more recent (67 generations–500 ybp) than in original analyses (200 generations–1,500 ybp).
Model checking based on comparisons of test quantities between observed data and 100 pseudo-observed datasets generated using parameter values drawn from posterior distributions. (A) Analyses on the full dataset, (B) Analyses on a pruned dataset with misclassified wild individuals and individuals with a recent admixed ancestry removed. (C) Analyses on the full dataset, assuming that admixture between ancestral Malus domestica and M. sylvestris was more recent (67 generations–500 ybp) than in (A).
Genetic differentiation (FST) between cultivars of different geographic origins (N = 266). Cultivars of unknown origin have been removed.
Description of the Multiplex PCRs (MP01, MP02, MP03, MP04) used for microsatellite amplification.
Method used for approximate Bayesian computations on alternative datasets/admixture times.
Acknowledgments
We thank Jacqui A. Shykoff, Rémy Petit, Jérome Enjalbert, Domenica Manicacci, Thierry Robert, Gwendal Restoux, and three anonymous reviewers for helpful suggestions and comments. We thank Plateforme de Génotypage GENTYANE INRA UMR 1095 Génétique Diversité et Ecophysiologie des Céréales Pauline Lasserre for help wit Aurélien Tellier, Virginie Ravigné, Peter Beerli, and Daniel Wegmann for help with, and advice on, data analysis. We thank Eric van de Weg for assistance in the determination of cultivar pedigrees. We thank the following for sampling and providing access to samples: Catherine Peix, Aymar Dzhangaliev, and collaborators in Kazakhstan, Evelyne Heyer (Museum National d'Histoire Naturelle, France), Marie-Anne Félix (Institut Jacques Monod, France), and Emmanuelle Jousselin (Centre de Biologie et de Gestion des Populations, France) for M. sieversii Dominique Beauvais (Abbaye de Beauport, Paimpol, France) and Jean Pierre Roullaud (Verger Conservatoire d'Arzano, France) for providing M. domestica Ara Hovhannisyan, Karen Manvelyan, and Eleonora Gabrielian for M. orientalis Alberto Dominici and Emanuela Fabrizi (Monti Simbruini Regional Park, Italy), Jan Kowalczyk and Dzmitry Kahan (Forest Research Institute, Poland), J?rg Kleinschmit and Wilfried Stei

我要回帖

更多关于 jounery into mystery 的文章

 

随机推荐