Skip to main content
Ecology and Evolution logoLink to Ecology and Evolution
. 2019 Dec 20;10(1):292–306. doi: 10.1002/ece3.5894

Association between blooming time and climatic adaptation in Prunus mume

Ting Shi 1,2, Wenjie Luo 1, Hantao Li 1, Xiao Huang 1, Zhaojun Ni 1, Haidong Gao 3, Shahid Iqbal 1, Zhihong Gao 1,
PMCID: PMC6972806  PMID: 31988729

Abstract

Prunus mume Sieb. et Zucc. is an important fruit crop of the subtropical region, originating in China. It blooms earlier than other deciduous fruit trees, but different regions have different blooming periods. The time of anthesis is related to the dormancy period, and a certain amount of chilling promotes bud break and blooming. To identify the relationship between blooming time and the climatic adaptation of P. mume cultivars in China, the nuclear and chloroplast genomes of 19 cultivars from the main cultivation areas of P. mume in China were resequenced. The average depth of coverage was 34X–76X, and a total of 388,134 single nucleotide polymorphisms were located within the coding regions of the gene (CDs). Additionally, the 19 cultivar accessions were divided into three groups based on their blooming time: early, mid, and late. Associated with the blooming time groups, 21 selective sweep regions were identified, which could provide evidence supporting the possible model of P. mume domestication originating due to natural selection. Furthermore, we identified a flowering gene, FRIGIDA‐LIKE 3 (FRL3), seems to affect the blooming time and the climatic adaptation of P. mume cultivars. This study is a major step toward understanding the climatic adaptation of P. mume cultivars in China.

Keywords: blooming time, climatic adaptation, genome resequencing, population genetics, Prunus mume


Twenty‐one selective sweep regions were identified by genome resequencing from 19 Prunus mume accessions, which could provide evidence supporting the model of P. mume domestication originating due to natural selection. Furthermore, we identified a flowering gene, FRIGIDA‐LIKE 3 (FRL3), seems to affect the chilling requirements and the climate adaptation of P. mume cultivars. This study is a major step toward understanding the climate adaptation of P. mume cultivars in China.

graphic file with name ECE3-10-292-g008.jpg

1. INTRODUCTION

Prunus mume Sieb. et Zucc. is a member of the Rosaceae family, a culturally important deciduous fruit tree in east Asia, including Japan, Korea, and southern China (Chu, 1999). It is mainly used for making liqueurs, pickles, and sauces (Adachi et al., 2007). P. mume originates in China and has been domesticated for thousands of years (Chu, 1999; Li, Chen, & Zhang, 2007). The area of Hengduan Mountain is considered the center of natural distribution and genetic diversity of P. mume (Chu, 1999). Moreover, the wild species were found in south China such as Guangdong and Fujian (Chu, 1999; Li, 2010).

Prunus mume blooms earliest in the spring, with a shorter dormancy period among all deciduous fruit trees. The seasonal dormancy is crucial for winter survival and regulation of timely growth activity of perennial plants in temperate and cold climates (Arora, Rowland, & Tanino, 2003). Bud dormancy has been classified into three categories: paradormancy, endodormancy, and ecodormancy (Cooke, Eriksson, & Junttila, 2012; Faust, Liu, Millard, & Stutte, 1991). Chilling requirements affects bud break and flowering because of its stimulation toward dormancy. P. mume has a wide range of chilling requirements from 479 CU to 1,323 CU corresponding with cultivars from Southern China to Northern China (Gao et al., 2012). The chilling requirements of the varieties in southern China such as Guangdong and Fujian province are early while that of varieties in central China such as Jiangsu and Zhejiang province are late. The varied blooming time of varieties may be due to different climatic conditions in different regions (Zhuang, Shi, Gao, Zhang, & Zhang, 2013).

DORMANCY‐ASSOCIATED MADS box6 (PmDAM6), a MADS box gene, was a candidate for growth inhibition and dormancy‐controlling gene (Sasaki et al., 2011). PmDAM3, PmDAM5, and PmDAM6 expressions are closely associated with dormancy release in both flower and vegetative buds. PmDAM1 and PmDAM6 could form heteromeric complexes with C‐repeat binding factor 5 (CBF5), a cold response signal factor. PmCBF1, PmCBF3, and PmDAM4 recognized the promoter of PmDAM6 by the alternative binding sites. Moreover, PmDAM6 could interact with a homolog of SUPPRESSOR OF OVEREXPRESSION OF CONSTANS1 (PmSOC1; Kitamura, Takeuchi, Yamane, & Tao, 2016). The interactions of these genes were key factors for early dormancy release and blooming time in P. mume, which made it sensitive to temperature changes, resulting in short dormancy and early flowering (Kitamura et al., 2016; Zhao et al., 2018). Other than these regulators, there are many other genes have been reported to be punitively involved in dormancy release, such as RGL2 (Lv, Huo, Wen, Gao, & Khalil‐Ur‐Rehman, 2018) and hormone‐related genes (Wen et al., 2016). PmRGL2 (REPRESSOR‐OF‐GA‐LIKE2), a DELLA protein, could play a negative role in bud dormancy release by regulating the GA biosynthetic enzymes (Lv et al., 2018). Gibberellin (GA 4) has a significant effect on promoting dormancy release in flower buds of P. mume (Zhuang, Gao, et al., 2013; Zhuang et al., 2015). Abscisic acid (ABA) is another plant hormone involved in regulating the onset of dormancy and in maintaining the dormant state (Karssen, Brinkhorst‐van der Swan, Breekland, & Koornneef, 1983; Wen et al., 2016). Furthermore, the QTL analyses localized the significant QTLs controlling leaf bud chilling requirements and heat requirements, leafing date, and PmDAM6 expression in leaf buds to a region in linkage group LG4, which suggests that this locus controls dormancy release, bud break, and PmDAM6's downregulation in P. mume leaf buds (Kitamura et al., 2018). QTLs for chilling requirements and blooming time were identified on the LG4 of almond (Sánchez‐Pérez, Dicenta, & Martínez‐Gómez, 2012) and sweet cherry (Castède et al., 2014). The QTLs for apricot blooming initiation time were located on LG1 and LG4 (Dirlewanger et al., 2012; Olukolu et al., 2009; Socquet‐Juglard et al., 2013). In peach, major QTLs for chilling requirements and blooming time were detected on LG1, LG4, and LG7 (Bielenberg et al., 2015; Fan et al., 2010; Zhebentyayeva et al., 2014). Genome‐wide association studies for CR in peach identified seven association peaks, located on LG 1, 3, 7, and 8. The strong association peak on LG 1 overlapped with a known major CR QTL (qCR1a) and co‐localized with the EVG locus underlying the evergrowing peach dormancy mutation (evg). DAM1‐6 were identified in qCR1a conferring a nondormancy phenotype in evg peach (Li et al., 2019). There were three most significant QTLs associated with 2.8, 1.8, and 1.0 days bloom delay, respectively, in sour cherry (Prunus cerasus L.). These QTLS were also demonstrated to have additive effects on delaying blooming date for both individual and multiple QTLs (Cai et al., 2018).

Next‐generation sequencing (NGS) technologies are faster and cheaper than Sanger sequencing, which is frequently used in plant studies and allows for a deeper genome variant analysis (Deschamps & Campbell, 2010; Jackson, Iwata, Lee, Schmutz, & Shoemaker, 2011). So far, genome sequencing and evolutionary genetics have provided information about the origin, evolution (Ellegren, 2014; Sedivy, Wu, & Hanzawa, 2017; Velasco, Hough, Aradhya, & Ross‐Ibarra, 2016; Wu, Terol, et al., 2018; Yu et al., 2018), and domestication (Akagi, Hanada, Yaegaki, Gradziel, & Tao, 2016; Myles et al., 2011; Qiu et al., 2015; Velasco et al., 2016; Wu, Wang, et al., 2018). The first 237M long genomic map of P. mume was constructed in 2012, and the actual size of the genome of P. mume was estimated to be about 280M. The processes of chromosome fusion and fragmentation in three Genera of Malus, Fragaria, and Prunus were analyzed (Zhang et al., 2012). Furthermore, Zhang et al constructed the phylogenetic tree of P. mume cultivars by using related species of Prunus armeniaca and Prunus persica as reference. After resequencing and genome assembly with P. armeniaca and P. persica and in conjunction with the published genomes of P. persica and P. mume, GWAS analysis identified multiple quantitative trait locus regions and found that the MYB108 gene was associated with flower color (Zhang et al., 2018).

To identify the climatic adaptation and relationship among the domesticated P. mume in China, 19 cultivated P. mume accessions from Yunnan, Sichuan, Guizhou, Jiangsu, Zhejiang, Hunan, Guangdong, Fujian, and the Taiwan provinces, the main cultivation area of P. mume in China, were resequenced and their accessions were locally grown for many years to be strong representatives of the cultivated P. mume. Based on analyses of population genetics and the evolution of cultivated P. mume, and selective sweeps associated with the analysis of the blooming time of the related genes, this study aimed to propose a model to explain the relationship between the climatic adaptation and blooming time of cultivated P. mume.

2. MATERIALS AND METHODS

2.1. Sampling information and genome resequencing

In this present study, a total of 19 P. mume accessions were collected and resequenced. All the accessions were from nine provinces in China (Jiangsu: R1 and R2; Sichuan: R3, R4, and R5; Yunnan: R06 and R07; Fujian: R08 and R09; Guangdong: R10 and R11; Guizhou: R12; Taiwan: R13 and R14; Zhejiang: R15, R16 and R17; and Hunan: R18 and R19, see Table 1). These sampling areas were divided into four parts: Jiangsu and Zhejiang belonged to east China; Hunan belonged to central China; Fujian, Guangdong and Taiwan belonged to south China; and Sichuan, Yunnan, and Guizhou belonged to southeast China. Genomic DNA was extracted from leaves using the CTAB method. Paired‐end DNA libraries (TruSeq® DNA Library Prep Kits; Illumina) with short inserts (~500 bp) were constructed according to the manufacturer's instructions and sequenced using the HiSeq™ ×10 platforms (Illumina).

Table 1.

Information of the blooming time partition of 19 Prunus mume varieties

Blooming time Cultivar Cultivar no. Region Full‐blossom period Weather station number and location
EARLY Taiwandaqingmei R13 Nantou, Taiwan Early February 59,316 Shantou
Taiwanyanzhimei R14 Nantou, Taiwan Mid‐January 59,316 Shantou
Ruanzhidalimei R10 Puning, Guangdong Mid‐January 59,316 Shantou
Shuangshuidaroumei R11 Puning, Guangdong Mid‐January 59,316 Shantou
Longyanmei R08 Yongtai, Fujian Mid‐January 58,847 Fuzhou
Baifenmei R09 Zhaoan, Fujian Early January 59,316 Shantou
MID Yunnanyanmei R06 Lijiang, Yunnan Mid‐January 56,651 Lijiang
Yunnanzhaoshuimei R07 Lijiang, Yunnan Mid‐January 56,651 Lijiang
Guizhousuanmei R12 Libo, Guizhou Early February 59,023 Hechi
Tonglvmei R18 Changsha, Hunan Late February 57,662 Changde
Xianmimei R19 Changsha, Hunan Late February 57,662 Changde
LATE Sichuanqingmei R03 Dayi, Sichuan Late February 56,394 Chengdu
Sichuanbaimei R04 Dayi, Sichuan Late February 56,394 Chengdu
Sichuanhuangmei R05 Dayi, Sichuan Late February 56,394 Chengdu
Nanhongmei R01 Nanjing, Jiangsu Early March 58,238 Nanjing
Hongguangmei R02 Nanjing, Jiangsu Early March 58,238 Nanjing
Ruantiaohongmei R15 Chaoshan, Zhejiang Mid‐March 58,454 Hangzhou
Xiaoyezhugan R16 Chaoshan, Zhejiang Early March 58,454 Hangzhou
Qingjia No.2 R17 Chaoshan, Zhejiang Mid‐March 58,454 Hangzhou

2.2. Investigation of the blooming time of Prunus mume accessions

Because of the local climatic conditions, such as temperature and humidity are very important factors for the blooming time and the chilling hours is one of the most important conditions for the adaptation of P. mume. According to the phenological period of each accession and the temperature data of each meteorological station over the past 60 years from the National Centers for Environmental Information (NCEI; gis.ncdc.noaa.gov), we roughly investigate the blooming time of these 19 varieties of P. mume used in this study.

2.3. Chloroplast genome resequencing

Chloroplast DNA was extracted from the fresh leaves of each accession using a modified CTAB method. A DNA concentration >50 ng/µl was measured using a NanoDrop spectrophotometer and fragmented through sonication. Then, fragmented DNA was purified and end‐repaired, and the size was determined through gel electrophoresis. The PCR products were constructed as short‐insert (300 bp) libraries using Illumina Nextera XT and then sequenced using the HiSeq ×10 platform (Illumina). Raw reads were filtered using NGSQC Tool kit v2.3.3 to obtain high‐quality reads, and then, the chloroplast genomes were assembled by NOVOPlasty using clean data and annotated with CpGAVAS (Hu et al., 2016). The chloroplast genome of P. persica (NCBI Accession number NC_014697.1) was used as the reference genome in this study. The MIcroSAtellite identification tool (MISA) was used to analyze the SSR (simple sequence repeat) from the chloroplast genome.

2.4. Genome mapping and variant calling

Bwa software (Li & Durbin, 2009) was used to align the reads obtained from resequencing onto the P. mume reference genome (Zhang et al., 2012), and subsequent mutation analysis was performed. Based on the position of clean reads on the reference genome, the sequencing depth, genomic coverage, and other information from each sample were measured and the mutation was detected. Single nucleotide polymorphism (SNP) detection was primarily implemented using GATK (McKenna et al., 2010). Based on the location results of clean reads in the reference genome, using SAMtools (Li et al., 2009) to mark duplicates, GATK was used for local realignment, base recalibration, and other pretreatments to ensure the detection of SNP accuracy and then use GATK for SNP and Indel detection, filtering, and obtaining the final SNP loci. SNP and small Indel annotation were performed using SnpEff (Cingolani et al., 2012), which predicted the effects of mutations. The structural variation (SV) was detected by BreakDancer (Fan, Abbott, Larson, & Chen, 2014). There are 5 types of structural variations: deletion (DEL), insertion (INS), and inversion (INV) intra chromosomal translocation (ITX), and interchromosomal translocation (CTX). The results of the various types of detected mutations are shown in the circos diagram, using the circos software, http://circos.ca/. All the variant genes were annotated using BLASTx in the nonredundant protein (NR), Swiss‐Prot, Clusters of Orthologous Groups (COG), Gene Ontology (GO) database, and Kyoto Encyclopedia of Genes and Genomes (KEGG), and these annotations were performed according to the method of previous reports (Conesa & Gotz, 2008). The following population genetic summary statistics were then calculated for each dataset: nucleotide diversity (π; Tajima, 1989), heterozygosity (Hetobs), the inbreeding coefficient (F IS), and differentiation, as given by F ST (Zhang et al., 2018). Nucleotide diversity is related to expected heterozygosity and is an overall measure of genetic variation.

2.5. Evolution analysis

To acquire more intuitive information of the evolution analysis, a principal components analysis (PCA) was performed using GCTA software, and the main clustering element was obtained. Correlation coefficients (r 2) of the alleles of the 19 varieties were calculated using PLINK software with a pair‐wise algorithm to measure the linkage disequilibrium (LD) value.

2.6. Population genetics analysis, identification of selective sweeps, and gene flow analysis

ADMIXTURE was used to infer population structure among the 19 samples, based on the SNPs. The population groups (K) varied from 2 to 10. Based on the SNPs of 19 accessions, using the peach as an exogenous reference species, a phylogenetic tree was constructed and PCA analysis was performed. The phylogenetic analysis was performed using FastTree software with the neighbor‐joining method.

To identify the genomic regions that may have been subjected to selection during domestication, we calculated nucleotide diversity (π), genetic differentiation (F ST), and the Watterson estimator (θ π) using a sliding‐window approach (50‐kb windows sliding in 10‐kb steps) with SNPs of the whole genome. The log2θ πratio was calculated as the nucleotide diversity values for the cultivars with late‐blooming and early blooming time. The regions with significantly late F ST values (in the 5% right tail of the empirical distribution of F ST values) and a significant reduction in diversity (in the 5% right tail of the empirical distribution of the log2θ πratio) were considered candidate selective sweeps.

Genes in the candidate selective sweeps were grouped into 1‐Mb nonoverlapping segments throughout the genome, according to their physical locations. Adjacent segments and segments with no more than 1‐Mb were separated and merged, while segments containing no more than five genes were discarded. GO term enrichment analyses were conducted for the genes in each merged segment along the genome using GO::TermFinder.

To estimate the gene flow between the high, mid, and early blooming time of P. mume, we used Treemix version 1.13 to investigate the gene flow between groups/subgroups.

2.7. Blooming time‐related genes analysis

Phenotypic data of blooming time were used for the association study. For the grouping of blooming time (Table 1), “low,” “mid,” and “high” groups were used. The final set of SNPs were filtered to keep those with the missing rate ≤10%, MAF ≥5%, SNP calling rate per accession ≥80%, and Hardy–Weinberg equilibrium test p value ≤.001. The resulting SNPs were used to perform GWAS for the blooming time using a linear mixed model (LMM) implemented in the FaST‐LMM program (Martinez et al., 2018). Raw p‐values were adjusted for multiple testing using the Benjamini–Hochberg procedure (Benjamini & Hochberg, 1997), and a significant association was based on a false discovery rate threshold of 0.01.

3. RESULTS

3.1. Blooming time and climate warming in the cultivated area of Prunus mume

Based on flowering time and the hours at 0–7.2°C, we divided the 19 varieties into three groups of blooming time: early, mid, and late (Table 1). The data for hours at 0–7.2°C in R01 and R02 were from the observation stations of Nanjing; R03, R04, and R05 were from Chengdu; R06 and R07 were from Lijiang; R08 and R09 were from Fuzhou; R10, R11, R13, and R14 were from Shantou; R12 was from Libo; R15, R16, and R17 were from Hangzhou; and R18 and R19 were from Changde (Table 1).

3.2. Genome, chloroplast genome resequencing, and genetic analysis of Prunus mume

Genomes of the 19 cultivated P. mume accessions were resequenced using Illumina HiSeq ×10. A total of 239.33 GB of data were obtained, and the Q30 was over 82.24%, with an average mapping rate of 93.9% (Additional file: Tables S1 and S2). The average depth of coverage was 34X–76X, and the genome coverage was 87.86% (at least one base was covered; Additional file: Table S3).

A total of 38,321,207 SNPs were identified (Additional file: Table S4). A total of 388,134 SNPs were located within the coding regions of the gene (CDs). We also identified 71,265 INDEL and 36,659 SV in the CDs (Table 2). They are filtered to only those that have a potential functional impact.

Table 2.

Classification and statistics of nucleotide diversity produced by various mutations

Cultivar no. SNP INDEL SV Total affected genes
R01 18,343 3,255 0 18,578
R02 19,226 3,498 1,772 19,606
R15 20,616 3,674 3,864 21,045
R16 20,709 3,717 3,243 21,092
R17 20,739 3,729 3,839 21,150
R18 20,754 3,892 1,539 21,044
R19 20,713 3,859 1,580 20,957
R08 20,631 3,779 1,723 20,932
R09 20,966 3,950 1,645 21,222
R10 20,741 3,821 1,734 21,014
R11 20,710 3,818 1,761 21,005
R13 20,578 3,733 1,733 20,870
R14 20,790 3,894 1,835 21,063
R03 20,842 3,968 1,719 21,107
R04 20,514 3,759 1,792 20,819
R05 20,556 3,790 1,768 20,856
R06 20,682 3,844 1,814 20,980
R07 20,625 3,806 1,681 20,897
R12 19,399 3,479 1,617 19,739

Abbreviations: SNP, single nucleotide polymorphism; SV, structural variation.

The average number of nucleotide diversities produced by these mutations was 20,739 among these 19 cultivated accessions (Table 2), and the summary of genetic diversity in P. mume cultivars is shown in Table 3.

Table 3.

Summary of genetic diversity in Prunus mume cultivars

Population Sites Polymorphic Invariant HetObs π F IS
R01 3,757,493 766,270 2,991,223 0.203931 0.101966 −1
R02 3,744,896 990,688 2,754,208 0.264544 0.132272 −1
R03 3,707,982 969,817 2,738,165 0.261548 0.130774 −1
R04 3,696,447 710,257 2,986,190 0.192146 0.096073 −1
R05 3,696,432 752,846 2,943,586 0.203668 0.101834 −1
R06 3,702,222 835,330 2,866,892 0.225629 0.112815 −1
R07 3,703,072 817,899 2,885,173 0.22087 0.110435 −1
R08 3,695,026 772,295 2,922,731 0.209009 0.104505 −1
R09 3,724,660 999,430 2,725,230 0.268328 0.134164 −1
R10 3,689,825 815,105 2,874,720 0.220906 0.110453 −1
R11 3,696,730 810,340 2,886,390 0.219205 0.109602 −1
R12 3,717,058 779,507 2,937,551 0.209711 0.104855 −1
R13 3,684,153 719,034 2,965,119 0.195169 0.097585 −1
R14 3,709,958 916,329 2,793,629 0.246992 0.123496 −1
R15 3,580,820 717,627 2,863,193 0.200409 0.100204 −1
R16 3,581,599 735,908 2,845,691 0.205469 0.102735 −1
R17 3,625,457 833,616 2,791,841 0.229934 0.114967 −1
R18 3,688,694 840,950 2,847,744 0.22798 0.11399 −1
R19 3,678,202 818,311 2,859,891 0.222476 0.111238 −1

Abbreviations: F IS, inbreeding coefficient; Hetobs, heterozygosity; π, nucleotide diversity.

A summary of the population genetic statistics is presented in Table 3. The lowest observed heterozygosity (Hetobs) in the accessions was 0.192 (Sichuan cultivar: R04), and the highest Hetobs was 0.268 (Fujian cultivar: R09). The genetic diversity (π) of cultivated P. mume was estimated to range from 0.096 to 0.134 (Table 3).

3.3. Population phylogenetic analysis of Prunus mume cultivars

A subset of 388,134 SNPs was screened in greater detail to construct a neighbor‐joining tree, using the peach genome as an outgroup. The 19 cultivars were clustered into two independent clades after branching from the peach (Figure 1a). The red cluster contained four cultivars with late‐blooming time (Jiangsu and Sichuan cultivars), four cultivars with mid‐blooming time (Yunnan, Guizhou and one cultivar in Hunan), and two cultivars with early blooming time (Fujian cultivars). For the green cluster, there were four cultivars with early blooming time (Taiwan and Guangdong cultivars), one Hunan cultivar with a mid‐blooming time, and three cultivars with late‐blooming time (Zhejiang cultivars).

Figure 1.

Figure 1

Phylogenetic trees of 19 Prunus mume cultivars: (a) Genome phylogenetic tree; (b) chloroplast phylogenetic tree

To get a better understanding of the phylogenetic relationships of P. mume, we assembled and analyzed the chloroplast genome of P. mume. Chloroplast genome resequencing of the 19 accessions was completed, and 239.3 GB of clean data was obtained, in which the percentage of the Q30 base was more than 82.24%. The chloroplast assembly results from 17 samples were very good, and they were able to produce a full‐circle figure with zero gaps. In addition, two samples of R16 and R17 could not be assembled into a circle, but could form better scaffolds (Additional file: Figure S1). Furthermore, the phylogenetic tree of the chloroplast showed the same cluster pattern as the genome phylogenetic tree (Figure 1b).

We used the program ADMIXTURE to fit a model of admixture, in which an individual's genome is composed of sites from up to K ancestral populations. We explored K = 1 through six ancestral populations (Figure 2) to investigate how assumptions regarding K impact the inference of population structure. Assuming a K = 2 admixture model, population admixture patterns were driven by geographic location (Zhejiang: R15, R16 and R17; Hunan: R19; Guangdong: R10 and R11, and Taiwan: R13 and R14; Figure 2a, top panel). However, higher Ks showed substantial substructure in all three blooming time populations. Log likelihoods for successively increasing the levels of K continued to increase substantially as K increased (Figure 2b), which was not unexpected since higher values of K add more parameters to the model, thereby improving the fit.

Figure 2.

Figure 2

Population structure of cultivated Prunus mume accessions. (a) ADMIXTURE clustering analysis using a high‐density dataset. K = 1–6 clusters are shown. (b) The estimated prediction error was minimum on a grid of K = 2, which suggests that it is the most suitable K. (c) The admixed genetic component and geographic distribution of P. mume accessions (K = 2). The size of the circle represents the blooming time, where a small circle represents earlier blooming time, and a bigger circle represents later blooming time

Linkage disequilibrium (LD) analysis showed that genomes have relatively short LD distances and relatively rapid LD decays (Figure 3a). The average r 2 value among P. mume SNPs, corresponding to the LD levels of the population, was relatively low. The average distance over which LD decayed to ~50% of its maximum value in cultivated P. mume was very short. Figure 3b–d showed the distribution of each accession in PCA space.

Figure 3.

Figure 3

Decay of (a) linkage disequilibrium (LD) and (b, c, d) principal component analysis (PCA)

3.4. Natural selection of Prunus mume in the process of domestication

To further study the genes in the transfer process, we performed selective clearance analysis. Because blooming time between the EARLY and LATE groups was significantly different, we used the EARLY and LATE‐blooming time groups to perform the selective sweep analysis (Figure 4a). Twenty‐one selective sweep regions and a total of 283 genes had been found in 19 selective sweep regions (Table S5). Considering the climate warming, the remarkable role of blooming time in distinguishing P. mume cultivars, genomic regions that were dramatically affected by selection during P. mume domestication were identified between the late‐blooming and early blooming time cultivars. The selected regions of early blooming time cultivars had a mean size of 173.8 kb, covered a total length of 3.65 Mb (1.3% of the genome; Table 4, and harbored 371 genes (Figure 4b, Table S5). The identified selective sweeps were enriched with genes associated with post‐translational modification, protein turnover, chaperones, replication, recombination and repair, transcription, translation, ribosomal structure and biogenesis, cell cycle control, cell division, chromosome partitioning, RNA processing and modification, chromatin structure and dynamics, cytoskeleton, energy production and conversion, amino acid transport and metabolism, nucleotide transport and metabolism, carbohydrate transport and metabolism, lipid transport and metabolism, inorganic ion transport and metabolism, signal transduction mechanisms, intracellular trafficking, secretion and vesicular transport, secondary metabolites biosynthesis, transport, and catabolism (Figure 4a). These biological processes are usually involved in natural selection during the process of domestication. According to the gene flow analysis, no gene flow occurred between the late and early cultivars (Figure 4c).

Figure 4.

Figure 4

Selective sweeps and gene flow. (a) Selective sweeps and QTLs related to transcription, translation, post‐translational modification, and secondary metabolism in Prunus mume. The lines of each linkage group indicate selective sweeps; the colorful box, triangle and circle of each linkage group indicate QTLs. A total of 21 selective sweeps in early blooming time cultivars showed coincidence with QTLs related to transcription, translation, post‐translational modification, and secondary metabolism. The numbering at the top of the chromosome is LG numbering in Dirlewanger' paper (Dirlewanger et al., 1998), the numbering below the chromosome is LG numbering in Zhang's paper (Zhang et al., 2012). (b) Distribution of log2θ πratio (High/Low) and F st of 50‐kb windows with 10‐kb steps. (c) Detection of gene flow among early, mid, and late‐blooming time Prunus mume cultivars

Table 4.

The location of the selective sweep regions between the groups with early and late‐blooming time

Chromosomes Starting site Ending site Region length (bp)
LG1 (LG6) 220,001 600,000 380,000
LG1 (LG6) 14,220,001 14,330,000 110,000
LG1 (LG6) 8,850,001 9,010,000 160,000
LG2 (LG1) 19,350,001 19,480,000 130,000
LG2 (LG1) 2,760,001 2,880,000 120,000
LG3 (LG4) 7,500,001 7,670,000 170,000
LG4 (LG3) 17,930,001 18,040,000 110,000
LG4 (LG3) 10,790,001 10,930,000 140,000
LG4 (LG3) 17,240,001 17,430,000 190,000
LG4 (LG3) 2,150,001 2,250,000 100,000
LG5 (LG2) 24,700,001 24,910,000 210,000
LG5 (LG2) 4,060,001 4,220,000 160,000
LG5 (LG2) 8,300,001 8,400,000 100,000
LG6 (LG8) 18,730,001 18,880,000 150,000
LG6 (LG8) 12,980,001 13,120,000 140,000
LG6 (LG8) 1,840,001 1,940,000 100,000
LG7 (LG5) 7,390,001 7,530,000 140,000
LG7 (LG5) 1,900,001 2,000,000 100,000
LG8 (LG7) 3,670,001 3,840,000 170,000
LG8 (LG7) 3,340,001 3,570,000 230,000
LG8 (LG7) 1,900,001 2,440,000 540,000

LG numberings in bracket indicate the LG numberings in the Dirlewanger's paper.

3.5. Genes related to blooming time for the climatic adaptation of cultivated Prunus mume

According to the blooming time groups, we identified a total of 127 SNPs, and 54 genes associated with these SNPs (SNPs within the gene or 2 kb upstream or downstream of the gene; Figure 5, Table S6). As it turns out, our method provided quite satisfactory results about GWAS, as shown by the p‐value of the Q‐Q plot (Figure S2), in which the postcorrection p‐values were found to be close to the expected curves. Through further analysis, FRL3 encoded FRIGIDA‐like protein 3, involved in cell differentiation and flowering, and was found on the LG1 (LG6 in Dirlewanger' paper). The relationship between SNPs related to FRL3 and blooming time in 19 cultivars was shown in the Figure 6.

Figure 5.

Figure 5

An illustration of a Manhattan plot depicting several single nucleotide polymorphisms that were strongly associated with blooming time

Figure 6.

Figure 6

Schematic illustration of the DNA regions in FRL3 alleles. The single nucleotide polymorphisms located at the FRL3 3′ UTR (6,400 bp) are indicated white boxes. Blue boxes represent FRL3 gene body. Early, mid, and late represent three types of blooming time

4. DISCUSSION

4.1. High genetic diversity in the fruiting of Prunus mume

Prunus is a large genus in the Rosaceae family, with more than 430 species, including multiple domesticated crops such as the almond (Prunus dulcis L.), apricot (P. armeniaca L.), cherry (Prunus avium L.), peach (P. persica L.), plum (Prunus domestica L.), and P. mume. Nucleotide diversity is related to expected heterozygosity and is an overall measure of genetic variation. F ST represents the amount of inbreeding due to random mating in a finite population and is used as a measure of population subdivision and genetic drift.

In this study, the lowest observed heterozygosity (Hetobs) in P. mume accessions was 0.192 (Sichuan cultivar: R04) and the highest was 0.268 (Fujian cultivar: R09), which is higher than the genetic diversity in plants and animals (Garcia‐Elfring, 2017). The genetic diversity of cultivated P. mume was estimated to range from 0.096 to 0.134. Zhang et al. resequenced 333 representative P. mume landraces and noted that their genetic diversity was 2.01 × 10−3 (Zhang et al., 2018), which is lower than that of our data. The main reason for this difference may be that the accessions Zhang et al. used were ornamental P. mume, but our accessions were fruiting P. mume. It was also indicated that natural selection played an essential role during the process of the domestication of fruiting P. mume, but for the ornamental P. mume, artificial selection might have been the main factor during domestication. Many ancient books in China also record the ornamental P. mume cultivars and planting modes, but there was scarcely any reference record of the planting modes of fruiting P. mume.

For peaches, a close relative of P. mume, Akagi et al. analyzed modern fruit (F) and modern ornamental (O‐A) cultivars, finding a significant reduction in genetic diversity of the O‐A cluster (Akagi et al., 2016). Compared with the closely related wild species Prunus kansuensis, Prunus mira, and Prunus davidiana, the peach has low levels of genetic diversity (International Peach Genome Initiative et al., 2013). For the domesticated almond, it retains more genetic diversity than any of the peach species. Many woody perennial crop species, including almonds (Velasco et al., 2016), grapes (Myles et al., 2011), and apples (Gross, Henk, Richards, Fazio, & Volk, 2014), lack domestication bottlenecks, but maintain much of their ancestral genetic diversity (Velasco et al., 2016). Moreover, in this study, the analysis of selective sweep regions and gene flow also indicated that the fruiting P. mume barely experienced artificial selection. Altogether, this suggests that fruiting P. mume has high genetic diversity, with less domestication and artificial selection.

4.2. Effects of a warming climate on the distribution of cultivated Prunus mume

Kitamura et al. (2018) localized the significant QTLs controlling the leaf bud blooming time to a region on linkage group 4 (equal to LG3 in our study); they thought that this locus controlled dormancy release in P. mume leaf buds (Kitamura et al., 2018). In our study, the sweep selected regions in LG3 involved energy production and conversion, post‐translational modification, protein turnover, chaperones, inorganic ion transport and metabolism, replication, recombination and repair, and signal transduction metabolism. Most genes in this region are structural genes, and they might be involved in the blooming time or other processes related to blooming time. There is a receptor gene in the selective sweep regions, ultraviolet‐B receptor UVR8, that may receive environmental signals (Fernandez, Tossi, Lamattina, & Cassia, 2016). The UV‐B‐specific signaling component acts as a UV‐B photoreceptor and plays a key role in establishing UV‐protective responses in plants. When introducing a variety to a new region, the light factor must be considered, as the light intensity and light quality may be different between regions (Vanhaelewyn et al., 2016). To survive, the plant must adapt to the new environment. After receiving signals from environment, some transcription factors are involved in this adaptation, like TGA2, which is a SA‐responsive and NPR1‐dependent transcription activator (Fan & Dong, 2002); transcription factor MYB36‐like orchestrates Casparian strip formation to offer the potential for improved water and nutrient use efficiencies and enhanced resistance to abiotic stresses (Kamiya et al., 2015). Transcription factor‐like protein DPB is involved in the regulation of the G1/S transition and increases the DNA binding activity of E2F proteins after hetero‐dimerization (Kosugi & Ohashi, 2002). Transcription factor Pur‐alpha 1 specifically binds the purine‐rich double‐stranded telomeric repeated sequence 5′‐AAACCCTAA‐3′ found in promoter telo boxes (Tremousaygue, Manevski, Bardet, Lescure, & Lescure, 2010) and the ethylene‐responsive transcription factor RAP2‐12. Finally, mostly the protein and enzyme genes, such as photosystem II core complex proteins psbY (Plochinger, Schwenkert, Sydow, Schroder, & Meurer, 2016), protein phytochrome kinase substrate 4 (PKS4) (Fankhauser & Christie, 2015), and light‐inducible protein CPRF2, participate in the light reaction (Monir & Zhu, 2018). Abscisic acid‐insensitive 5‐like protein 1 (Wang, Li, Mao, Li, & Jing, 2016), ethylene‐responsive transcription factor RAP2‐12 (Kosmacz et al., 2015), auxin‐induced protein AUX28‐like (Xie et al., 2018), and auxin‐induced protein 22D‐like (Han et al., 2017) take part in plant growth and development regulated by plant hormones. E3 ubiquitin‐protein ligase CIP8‐like (Wang et al., 2019), E3 ubiquitin‐protein ligase RNF144B (Michel, Swatek, Hospenthal, & Komander, 2017), and ERAD‐associated E3 ubiquitin‐protein ligase HRD1B‐like are associated with ubiquitination (Wang, Ye, Lencer, & Hansen, 2006). Serine/threonine‐protein kinase (Dudek et al., 1997), 1‐phosphatidylinositol‐3‐phosphate 5‐kinase FAB1B (Hirano, Sato, & behavior, 2011), G‐type lectin S‐receptor‐like serine/threonine‐protein kinase (Sun et al., 2013), phosphatidylinositol 4‐phosphate 5‐kinase α (Honda et al., 1999), and so on are all genes in the selective sweep region that might allow plants to better adapt to the environment and have potential value for future studies.

We also used GWAS to analyze the genes associated with blooming time, despite only having 19 accessions. Our method provided quite satisfactory results, as shown by the p‐value of the Q‐Q plot, in which the p‐values after correction were found to be close to the expected curves. From the GWAS results, a gene related to flowering time was identified, FRL3. FRL3 is a family member of FRIGIDA (FRI). FRI and flowering locus C (FLC) are the two main genes that confer the flowering traits in the winter annual Arabidopsis thaliana (Arabidopsis Interactome Mapping Consortium, 2011; Burn, Smyth, Peacock, & Dennis, 1993; Clarke & Dean, 1994; Lee, Bleecker, & Amasino, 1993). The FRIGIDA superfamily contains plant proteins that are similar to the FRIGIDA protein expressed in Arabidopsis, containing five distinct FRI subfamilies, including FRI (I), FRL1/2 (II), FRL3 (III), FRL4/4a/4b (IV), and FRL5 (V) (Risk, Laurie, Macknight, & Day, 2010). FRL1 and FRL2 have redundant roles in promoting FLC expression in the presence of FRI (Michaels, Bezerra, & Amasino, 2004); however, the roles of the other FRI subfamilies remain unclear. FRL3 may be a flowering repressor, and delayed floral timing may be regulated by gibberellic acid in bamboo (Liu, Zhu, Lin, & Ma, 2016). So, FRL3 may be an indispensable gene in the climatic adaptation process of P. mume, to control flowering time and normal breeding for the local climate. Therefore, FRL3 might play an important role in adapting to different local climates.

4.3. A proposed model for the climatic adaptation of cultivated Prunus mume

To explore the phylogenetic relationships of these 19 cultivated P. mume varieties, we performed a phylogenetic analysis of all identified SNPs, using the peach as an outgroup (Figure 1a). In the phylogenetic trees, the 19 accessions divided into two subgroups. In the first subgroup, R01 and R02, the varieties from Jiangsu province, were grouped. These two varieties are particularly close to the outgroup (Figure 1). It also seems that the Jiangsu varieties are relatively independent of the other varieties. The second subgroup was R10, R11, R13, R14, R15, R16, R17, and R19, encompassing the varieties from Guangdong, Guizhou, Taiwan, Zhejiang, and Hunan provinces. A principle component analysis (PCA) also supported our view. The Fujian, Sichuan, and Yunnan varieties clustered together. The varieties from Hunan province, R18 and R19, were separated in the phylogenetic trees, R18 clustered with R07, which are the varieties from Guizhou province and R19 clustered with R17, which is the variety from Zhejiang province. We suspect that this is due to the geographic location of Hunan (nearly in the middle of the P. mume cultivated region in China), so Hunan province may have been a hub in the course of the transfer of P. mume that linked western and eastern China.

After this, we used the Bayesian clustering program ADMIXTURE, changing K progressively from 2 to 10. When K = 2, all the cultivars separated clearly into two parts. One part included the varieties from the mainland and Fujian, while the other included the coastal region varieties. When K = 3, these varieties split into three parts. The cultivars of the coastal region, except for Fujian, still formed a group‐like structure similar to K = 2, but the mainland varieties divided into two parts: one part was a cultivar from Jiangsu and Zhejiang, and the other part was the remaining varieties. Based on the phylogenetic trees, PCA analysis, and population structure analysis, we assume that the transfer direction of cultural P. mume had three directions. The first direction is from Yunnan, to Sichuan, Guizhou and Hunan provinces, close to the origin area (southwest China) of P. mume. The second direction is a transfer to Fujian, Guangdong, and Taiwan provinces, located to the southeast from Hunan. The last direction is from Hunan to Jiangsu and Zhejiang provinces, in mid‐eastern China. In the PCA and ADMIXTURE analyses, the Sichuan and Fujian varieties were all clustered together, and Hunan province may be the hub in the transfer process of P. mume (Figure 7).

Figure 7.

Figure 7

Climatic adaptation process of cultivated Prunus mume

5. CONCLUSION

In this study, a genome variation mapping 19 P. mume accessions, collected from nine provinces in China, was performed. We found that fruiting P. mume has higher genetic diversity than ornamental P. mume. Associated with the blooming time data, 21 selective sweep regions, with a total of 283 genes involved in transcription, translation, post‐translational modification, and mostly metabolism, were identified. There was no gene flow between the late and early cultivars, which provides evidence to support a possible model of P. mume domestication that might include natural selection. Furthermore, we identified a total of 127 SNPs and 54 genes associated with the blooming time, and the flowering gene FRL3, which could affect the blooming time and the climatic adaptation of P. mume cultivars. This study is a major step toward understanding the climatic adaptation of P. mume cultivars in China and further contributes a molecular foundation for the origination and evolution of P. mume.

CONFLICT OF INTEREST

None of the authors have any competing interests in the manuscript.

AUTHOR CONTRIBUTIONS

TS, WL, HL, and XH performed the experiments. ZN prepared the materials; TS and WL wrote the manuscript; HG analyzed the bioinformatics data; SI modified the language; ZG designed the experiments.

Supporting information

 

 

 

 

ACKNOWLEDGMENTS

This research was supported by the National Key Research and development Program of China (2018YFD1000107), the National Natural Science Foundation of China (31772282, 31971703), Project for Crop Germplasm conservation, the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD), the Earmarked Fund for China Agriculture Research System (CARS‐19) and China Postdoctoral Science Foundation (2018M640497) in materials collection, data analysis, and experiment.

Shi T, Luo W, Li H, et al. Association between blooming time and climatic adaptation in Prunus mume . Ecol Evol. 2020;10:292–306. 10.1002/ece3.5894

DATA AVAILABILITY STATEMENT

The sequence data of P. mume genome resequencing involved in this study are being deposited in NCBI (SRA accession is PRJNA561464). All other relevant data (Tables S5 and S6, Figures S1 and S2) supporting the findings of the study are available in this article: Figshare https://doi.org/10.6084/m9.figshare.10298780.v3.

REFERENCES

  1. Adachi, M. , Suzuki, Y. , Mizuta, T. , Osawa, T. , Adachi, T. , Osaka, K. , … Masuda, K. (2007). The “Prunus mume Sieb. et Zucc” (Ume) is a rich natural source of novel anti‐cancer substance. International Journal of Food Properties, 10, 375–384. [Google Scholar]
  2. Akagi, T. , Hanada, T. , Yaegaki, H. , Gradziel, T. M. , & Tao, R. (2016). Genome‐wide view of genetic diversity reveals paths of selection and cultivar differentiation in peach domestication. DNA Research, 23, 271–282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Arabidopsis Interactome Mapping Consortium (2011). Evidence for network evolution in an Arabidopsis interactome map. Science, 333, 601–607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Arora, R. , Rowland, L. J. , & Tanino, K. (2003). Induction and release of bud dormancy in woody perennials: a science comes of age. HortScience, 38, 911–921. [Google Scholar]
  5. Benjamini, Y. , & Hochberg, Y. (1997). Multiple hypotheses testing with weights. Scandinavian Journal of Statistics, Theory and Applications, 24, 407–418. [Google Scholar]
  6. Bielenberg, D. G. , Rauh, B. , Fan, S. , Gasic, K. , Abbott, A. G. , Reighard, G. L. , … Wells, C. E. (2015). Genotyping by sequencing for SNP‐based linkage map construction and QTL analysis of chilling requirement and bloom date in peach [Prunus persica (L.) Batsch]. PLoS One, 10, e0139406 10.1371/journal.pone.0139406 [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Burn, J. E. , Smyth, D. R. , Peacock, W. J. , & Dennis, E. S. (1993). Genes conferring late flowering in Arabidopsis thaliana . Genetica, 90, 147–155. [Google Scholar]
  8. Cai, L. , Stegmeir, T. , Sebolt, A. , Zheng, C. , Bink, M. C. A. M. , & Iezzoni, A. (2018). Identification of bloom date QTLs and haplotype analysis in tetraploid sour cherry (Prunus cerasus). Tree Genetics & Genomes, 14, 22. [Google Scholar]
  9. Castède, S. , Campoy, J. A. , García, J. Q. , Le Dantec, L. , Lafargue, M. , Barreneche, T. , … Dirlewanger, E. (2014). Genetic determinism of phenological traits highly affected by climate change in Prunus avium: Flowering date dissected into chilling and heat requirements. New Phytologist, 202, 703–715. [DOI] [PubMed] [Google Scholar]
  10. Chu, M. (1999). Chinese fruit tree: Prunus mume. Beijing, China: China Forestry Publishing House; (In Chinese). [Google Scholar]
  11. Cingolani, P. , Platts, A. , le Wang, L. , Coon, M. , Nguyen, T. , Wang, L. , … Ruden, D. M. (2012). A program for annotating and predicting the effects of single nucleotide polymorphisms, SnpEff: SNPs in the genome of Drosophila melanogaster strain w1118; iso‐2; iso‐3. Fly, 6, 80–92. 10.4161/fly.19695 [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Clarke, J. H. , & Dean, C. (1994). Mapping FRI, a locus controlling flowering time and vernalization response in Arabidopsis thaliana . Molecular and General Genetics, 242, 81–89. [DOI] [PubMed] [Google Scholar]
  13. Conesa, A. , & Gotz, S. (2008). Blast2GO: A comprehensive suite for functional analysis in plant genomics. International Journal of Plant Genomics, 2008, 619832. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Cooke, J. E. K. , Eriksson, M. E. , & Junttila, O. (2012). The dynamic nature of bud dormancy in trees: Environmental control and molecular mechanisms. Plant, Cell and Environment, 35, 1707–1728. [DOI] [PubMed] [Google Scholar]
  15. Deschamps, S. , & Campbell, M. A. (2010). Utilization of next‐generation sequencing platforms in plant genomics and genetic variant discovery. Molecular Breeding, 25, 553–570. [Google Scholar]
  16. Dirlewanger, E. , Pronier, V. , Parvery, C. , Rothan, C. , Guye, A. , & Monet, R. (1998). Genetic linkage map of peach [Prunus persica (L.) Batsch] using morphological and molecular markers. Theoretical and Applied Genetics, 97, 888–895. [Google Scholar]
  17. Dirlewanger, E. , Quero‐García, J. , Le Dantec, L. , Lambert, P. , Ruiz, D. , Dondini, L. , … Arús, P. (2012). Comparison of the genetic determinism of two key phenological traits, flowering and maturity dates, in three Prunus species: Peach, apricot and sweet cherry. Heredity, 109, 280 10.1038/hdy.2012.38 [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Dudek, H. , Datta, S. R. , Franke, T. F. , Birnbaum, M. J. , Yao, R. , Cooper, G. M. , … Greenberg, M. E. (1997). Regulation of neuronal survival by the serine‐threonine protein kinase Akt. Science, 275, 661–665. [DOI] [PubMed] [Google Scholar]
  19. Ellegren, H. (2014). Genome sequencing and population genomics in non‐model organisms. Trends in Ecology & Evolution, 29, 51–63. [DOI] [PubMed] [Google Scholar]
  20. Fan, S. , Bielenberg, D. G. , Zhebentyayeva, T. N. , Reighard, G. L. , Okie, W. R. , Holland, D. , & Abbott, A. G. (2010). Mapping quantitative trait loci associated with chilling requirement, heat requirement and bloom date in peach (Prunus persica). New Phytologist, 185, 917–930. [DOI] [PubMed] [Google Scholar]
  21. Fan, W. H. , & Dong, X. N. (2002). In vivo interaction between NPR1 and transcription factor TGA2 leads to salicylic acid‐mediated gene activation in Arabidopsis. The Plant Cell, 14, 1377–1389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Fan, X. , Abbott, T. E. , Larson, D. , & Chen, K. (2014). BreakDancer: identification of genomic structural variation from paired‐end read mapping. Current Protocols in Bioinformatics, 45, 15.6.1–15.6.11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Fankhauser, C. , & Christie, J. M. (2015). Plant phototropic growth. Current Biology, 25, R384–R389. [DOI] [PubMed] [Google Scholar]
  24. Faust, M. , Liu, D. , Millard, M. M. , & Stutte, G. (1991). Bound versus free water in dormant apple buds—A theory for endodormancy. HortScience, 26, 887–890. [Google Scholar]
  25. Fernandez, M. B. , Tossi, V. , Lamattina, L. , & Cassia, R. (2016). A comprehensive phylogeny reveals functional conservation of the UV‐B photoreceptor UVR8 from green algae to higher plants. Frontiers in Plant Science, 7, 1698. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Gao, Z. H. , Zhuang, W. B. , Wang, L. J. , Shao, J. , Luo, X. Y. , Cai, B. H. , & Zhang, Z. (2012). Evaluation of chilling and heat requirements in Japanese apricot with three models. Hortscience, 47, 1826–1831. [Google Scholar]
  27. Garcia‐Elfring, A. (2017). Admixture on the northern front: Population genomics of the white‐footed mouse (Peromyscus leucopus) in Quebec and evidence of hybridization with the deer mouse (Peromyscus maniculatus). Montreal, QC: McGill University. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Gross, B. L. , Henk, A. D. , Richards, C. M. , Fazio, G. , & Volk, G. M. (2014). Genetic diversity in Malus× domestica (Rosaceae) through time in response to domestication. American Journal of Botany, 101, 1770–1779. [DOI] [PubMed] [Google Scholar]
  29. Han, Y. Q. , Gao, Y. M. , Shi, Y. , Du, J. D. , Zheng, D. F. , & Liu, G. F. (2017). Genome‐wide transcriptome profiling reveals the mechanism of the effects of uniconazole on root development in glycine max. Journal of Plant Biology, 60, 387–403. [Google Scholar]
  30. Hirano, T. , & Sato, M. H. (2011). Arabidopsis FAB1A/B is possibly involved in the recycling of auxin transporters. Plant Signaling & Behavior, 6, 583–585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Honda, A. , Nogami, M. , Yokozeki, T. , Yamazaki, M. , Nakamura, H. , Watanabe, H. , … Kanaho, Y. (1999). Phosphatidylinositol 4‐phosphate 5‐kinase α is a downstream effector of the small G protein ARF6 in membrane ruffle formation. Cell, 99, 521–532. 10.1016/S0092-8674(00)81540-8 [DOI] [PubMed] [Google Scholar]
  32. Hu, Y. H. , Woeste, K. E. , Dang, M. , Zhou, T. , Feng, X. J. , Zhao, G. F. , … Zhao, P. (2016). The complete chloroplast genome of common walnut (Juglans regia). Mitochondrial DNA Part B, 1, 189–190. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. International Peach Genome Initiative , Verde, I. , Abbott, A. G. , Scalabrin, S. , Jung, S. , Shu, S. , … Rokhsar, D. S. (2013). The high‐quality draft genome of peach (Prunus persica) identifies unique patterns of genetic diversity, domestication and genome evolution. Nature Genetics, 45, 487–494. [DOI] [PubMed] [Google Scholar]
  34. Jackson, S. A. , Iwata, A. , Lee, S. H. , Schmutz, J. , & Shoemaker, R. (2011). Sequencing crop genomes: Approaches and applications. New Phytologist, 191, 915–925. [DOI] [PubMed] [Google Scholar]
  35. Kamiya, T. , Borghi, M. , Wang, P. , Danku, J. M. , Kalmbach, L. , Hosmani, P. S. , … Salt, D. E. (2015). The MYB36 transcription factor orchestrates Casparian strip formation. Proceedings of the National Academy of Sciences of the United States of America, 112, 10533–10538. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Karssen, C. M. , Brinkhorst‐van der Swan, D. L. , Breekland, A. E. , & Koornneef, M. (1983). Induction of dormancy during seed development by endogenous abscisic acid: Studies on abscisic acid deficient genotypes of Arabidopsis thaliana (L.) Heynh. Planta, 157, 158–165. 10.1007/BF00393650 [DOI] [PubMed] [Google Scholar]
  37. Kitamura, Y. , Habu, T. , Yamane, H. , Nishiyama, S. , Kajita, K. , Sobue, T. , … Tao, R. (2018). Identification of QTLs controlling chilling and heat requirements for dormancy release and bud break in Japanese apricot (Prunus mume). Tree Genetics and Genomes, 14, 101. [Google Scholar]
  38. Kitamura, Y. , Takeuchi, T. , Yamane, H. , & Tao, R. (2016). Simultaneous down‐regulation of DORMANCY‐ASSOCIATED MADS‐box6 and SOC1 during dormancy release in Japanese apricot (Prunus mume) flower buds. The Journal of Horticultural Science and Biotechnology, 91, 476–482. [Google Scholar]
  39. Kosmacz, M. , Parlanti, S. , Schwarzländer, M. , Kragler, F. , Licausi, F. , & Van Dongen, J. T. (2015). The stability and nuclear localization of the transcription factor RAP 2.12 are dynamically regulated by oxygen concentration. Plant, Cell and Environment, 38, 1094–1103. [DOI] [PubMed] [Google Scholar]
  40. Kosugi, S. , & Ohashi, Y. (2002). Interaction of the Arabidopsis E2F and DP proteins confers their concomitant nuclear translocation and transactivation. Plant Physiology, 128, 833–843. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Lee, I. , Bleecker, A. , & Amasino, R. (1993). Analysis of naturally occurring late flowering in Arabidopsis thaliana . Molecular and General Genetics, 237, 171–176. [DOI] [PubMed] [Google Scholar]
  42. Li, H. , & Durbin, R. (2009). Fast and accurate short read alignment with Burrows‐Wheeler transform. Bioinformatics, 25, 1754–1760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Li, H. , Handsaker, B. , Wysoker, A. , Fennell, T. , Ruan, J. , Homer, N. , … Genome Project Data Processing Subgroup (2009). The sequence alignment/map format and SAMtools. Bioinformatics, 25, 2078–2079. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Li, Q. (2010). Studies on investigation of wild Mei germplasm resources of Sichuan, Yunnan, Tibet and Guizhou province as well as regional tests of hardy Mei cultivars in North China. PhD Thesis, Beijing Forestry University. [Google Scholar]
  45. Li, Q. , Chen, J. , & Zhang, Q. (2007). Carbonated stone fruit from underground at Peiligang historical site in Xinzheng, Henan Province. Journal of Beijing Forestry University, 29, 59–61. [Google Scholar]
  46. Li, Y. , Cao, K. , Zhu, G. , Fang, W. , Chen, C. , Wang, X. , … Wang, L. (2019). Genomic analyses of an extensive collection of wild and cultivated accessions provide new insights into peach breeding history. Genome Biology, 20, 36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Liu, S. N. , Zhu, L. F. , Lin, X. C. , & Ma, L. Y. (2016). Overexpression of the repressor gene PvFRI‐L from Phyllostachys violascens delays flowering time in transgenic Arabidopsis thaliana . Biologia Plantarum, 60, 401–409. [Google Scholar]
  48. Lv, L. , Huo, X. , Wen, L. , Gao, Z. , & Khalil‐Ur‐Rehman, M. (2018). Isolation and role of PmRGL2 in GA‐mediated floral bud dormancy release in Japanese apricot (Prunus mume Siebold et Zucc.). Frontiers in Plant Science, 9, 27 10.3389/fpls.2018.00027 [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Martinez, H. , Barrachina, S. , Castillo, M. , Quintana‐Ort, I. E. , Rambla de Argila, J. , Farre, X. , & Navarro, A. (2018). FaST‐LMM for two‐way epistasis tests on high‐performance clusters. Journal of Computational Biology, 25, 862–870. [DOI] [PubMed] [Google Scholar]
  50. McKenna, A. , Hanna, M. , Banks, E. , Sivachenko, A. , Cibulskis, K. , Kernytsky, A. , … DePristo, M. A. (2010). The genome analysis toolkit: A MapReduce framework for analyzing next‐generation DNA sequencing data. Genome Research, 20, 1297–1303. 10.1101/gr.107524.110 [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Michaels, S. D. , Bezerra, I. C. , & Amasino, R. M. (2004). FRIGIDA‐related genes are required for the winter‐annual habit in Arabidopsis. Proceedings of the National Academy of Sciences of the United States of America, 101, 3281–3285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Michel, M. A. , Swatek, K. N. , Hospenthal, M. K. , & Komander, D. (2017). Ubiquitin linkage‐specific affimers reveal insights into K6‐linked ubiquitin signaling. Molecular Cell, 68, 233–246.e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Monir, M. M. , & Zhu, J. (2018). Dominance and epistasis interactions revealed as important variants for leaf traits of maize NAM population. Frontiers in Plant Science, 9, 627. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Myles, S. , Boyko, A. R. , Owens, C. L. , Brown, P. J. , Grassi, F. , Aradhya, M. K. , … Buckler, E. S. (2011). Genetic structure and domestication history of the grape. Proceedings of the National Academy of Sciences of the United States of America, 108, 3530–3535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Olukolu, B. A. , Trainin, T. , Fan, S. , Kole, C. , Bielenberg, D. G. , Reighard, G. L. , … Holland, D. (2009). Genetic linkage mapping for molecular dissection of chilling requirement and budbreak in apricot (Prunus armeniaca L.). Genome, 52, 819–828. [DOI] [PubMed] [Google Scholar]
  56. Plochinger, M. , Schwenkert, S. , von Sydow, L. , Schroder, W. P. , & Meurer, J. (2016). Functional update of the auxiliary proteins PsbW, PsbY, HCF136, PsbN, TerC and ALB3 in maintenance and assembly of PSII. Frontiers in Plant Science, 7, 423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Qiu, Q. , Wang, L. , Wang, K. , Yang, Y. , Ma, T. , Wang, Z. , … Liu, J. (2015). Yak whole‐genome resequencing reveals domestication signatures and prehistoric population expansions. Nature Communications, 6, 10283. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Risk, J. M. , Laurie, R. E. , Macknight, R. C. , & Day, C. L. (2010). FRIGIDA and related proteins have a conserved central domain and family specific N‐ and C‐ terminal regions that are functionally important. Plant Molecular Biology, 73, 493–505. 10.1007/s11103-010-9635-2 [DOI] [PubMed] [Google Scholar]
  59. Sánchez‐Pérez, R. , Dicenta, F. , & Martínez‐Gómez, P. (2012). Inheritance of chilling and heat requirements for flowering in almond and QTL analysis. Tree Genetics & Genomes, 8, 379–389. [Google Scholar]
  60. Sasaki, R. , Yamane, H. , Ooka, T. , Jotatsu, H. , Kitamura, Y. , Akagi, T. , & Tao, R. (2011). Functional and expressional analyses of PmDAM genes associated with endodormancy in Japanese apricot. Plant Physiology, 157, 485–497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Sedivy, E. J. , Wu, F. , & Hanzawa, Y. (2017). Soybean domestication: The origin, genetic architecture and molecular bases. New Phytologist, 214, 539–553. [DOI] [PubMed] [Google Scholar]
  62. Socquet‐Juglard, D. , Christen, D. , Devènes, G. , Gessler, C. , Duffy, B. , & Patocchi, A. (2013). Mapping architectural, phenological, and fruit quality QTLs in apricot. Plant Molecular Biology Reporter, 31, 387–397. [Google Scholar]
  63. Sun, X. L. , Yu, Q. Y. , Tang, L. L. , Ji, W. , Bai, X. , Cai, H. , … Zhu, Y. M. (2013). GsSRK, a G‐type lectin S‐receptor‐like serine/threonine protein kinase, is a positive regulator of plant tolerance to salt stress. Journal of Plant Physiology, 170, 505–515. 10.1016/j.jplph.2012.11.017 [DOI] [PubMed] [Google Scholar]
  64. Tajima, F. (1989). Statistical method for testing the neutral mutation hypothesis by DNA polymorphism. Genetics, 123, 585–595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Tremousaygue, D. , Manevski, A. , Bardet, C. , Lescure, N. , & Lescure, B. (2010). Plant interstitial telomere motifs participate in the control of gene expression in root meristems. The Plant Journal, 20, 553–561. [DOI] [PubMed] [Google Scholar]
  66. Vanhaelewyn, L. , Schumacher, P. , Poelman, D. , Fankhauser, C. , Van Der Straeten, D. , & Vandenbussche, F. (2016). REPRESSOR OF ULTRAVIOLET‐B PHOTOMORPHOGENESIS function allows efficient phototropin mediated ultraviolet‐B phototropism in etiolated seedlings. Plant Science, 252, 215–221. [DOI] [PubMed] [Google Scholar]
  67. Velasco, D. , Hough, J. , Aradhya, M. , & Ross‐Ibarra, J. (2016). Evolutionary genomics of peach and almond domestication. G3 (Bethesda), 6, 3985–3993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Wang, J. , Li, Q. , Mao, X. , Li, A. , & Jing, R. (2016). Wheat transcription factor TaAREB3 participates in drought and freezing tolerances in Arabidopsis . International Journal of Biological Sciences, 12, 257–269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Wang, X. , Ye, Y. , Lencer, W. , & Hansen, T. H. (2006). The viral E3 ubiquitin ligase mK3 uses the Derlin/p97 endoplasmic reticulum‐associated degradation pathway to mediate down‐regulation of major histocompatibility complex class I proteins. Journal of Biological Chemistry, 281, 8636–8644. 10.1074/jbc.M513920200 [DOI] [PubMed] [Google Scholar]
  70. Wang, Z. , Song, S. , Sheng, S. , Tian, J. , Wu, R. , & Pang, X. (2019). Comparative transcriptome analysis identifies differentially expressed genes between normal and late‐blooming Siberian apricot. Journal of Forestry Research, 30, 2277–2288. [Google Scholar]
  71. Wen, L. H. , Zhong, W. J. , Huo, X. M. , Zhuang, W. B. , Ni, Z. J. , & Gao, Z. H. (2016). Expression analysis of ABA‐ and GA‐related genes during four stages of bud dormancy in Japanese apricot (Prunus mume Sieb. et Zucc). The Journal of Horticultural Science and Biotechnology, 91, 362–369. [Google Scholar]
  72. Wu, G. , Terol, J. , Ibanez, V. , Lopez‐Garcia, A. , Perez‐Roman, E. , Borreda, C. , … Talon, M. (2018). Genomics of the origin and evolution of Citrus. Nature, 554, 311–316. [DOI] [PubMed] [Google Scholar]
  73. Wu, J. , Wang, Y. , Xu, J. , Korban, S. S. , Fei, Z. , Tao, S. , … Zhang, S. (2018). Diversification and independent domestication of Asian and European pears. Genome Biology, 19, 77. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Xie, R. , Ge, T. , Zhang, J. , Pan, X. , Ma, Y. , Yi, S. , & Zheng, Y. (2018). The molecular events of IAA inhibiting citrus fruitlet abscission revealed by digital gene expression profiling. Plant Physiology and Biochemistry, 130, 192–204. [DOI] [PubMed] [Google Scholar]
  75. Yu, Y. , Fu, J. , Xu, Y. , Zhang, J. , Ren, F. , Zhao, H. , … Xie, H. (2018). Genome re‐sequencing reveals the evolutionary history of peach fruit edibility. Nature Communications, 9, 5404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Zhang, Q. , Chen, W. , Sun, L. , Zhao, F. , Huang, B. , Yang, W. , … Wang, J. (2012). The genome of Prunus mume . Nature Communications, 3, 1318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Zhang, Q. , Zhang, H. , Sun, L. , Fan, G. , Ye, M. , Jiang, L. , … Cheng, T. (2018). The genetic architecture of floral traits in the woody plant Prunus mume . Nature Communications, 9, 1702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Zhao, K. , Zhou, Y. , Ahmad, S. , Xu, Z. , Li, Y. , Yang, W. , … Zhang, Q. (2018). Comprehensive cloning of Prunus mume dormancy associated MADS‐Box genes and their response in flower bud development and dormancy. Frontiers in Plant Science, 9, 457. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Zhebentyayeva, T. N. , Fan, S. , Chandra, A. , Bielenberg, D. G. , Reighard, G. L. , Okie, W. R. , & Abbott, A. G. (2014). Dissection of chilling requirement and bloom date QTLs in peach using a whole genome sequencing of sibling trees from an F 2 mapping population. Tree Genetics & Genomes, 10, 35–51. [Google Scholar]
  80. Zhuang, W. , Gao, Z. , Wang, L. , Zhong, W. , Ni, Z. , & Zhang, Z. (2013). Comparative proteomic and transcriptomic approaches to address the active role of GA4 in Japanese apricot flower bud dormancy release. Journal of Experimental Botany, 64, 4953–4966. 10.1093/jxb/ert284 [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Zhuang, W. , Gao, Z. , Wen, L. , Huo, X. , Cai, B. , & Zhang, Z. (2015). Metabolic changes upon flower bud break in Japanese apricot are enhanced by exogenous GA4. Horticulture Research, 2, 15046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Zhuang, W. , Shi, T. , Gao, Z. , Zhang, Z. , & Zhang, J. (2013). Differential expression of proteins associated with seasonal bud dormancy at four critical stages in Japanese apricot. Plant Biology, 15, 233–242. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

 

 

 

 

Data Availability Statement

The sequence data of P. mume genome resequencing involved in this study are being deposited in NCBI (SRA accession is PRJNA561464). All other relevant data (Tables S5 and S6, Figures S1 and S2) supporting the findings of the study are available in this article: Figshare https://doi.org/10.6084/m9.figshare.10298780.v3.


Articles from Ecology and Evolution are provided here courtesy of Wiley

RESOURCES