Molecular phylogeny and inflorescence evolution of Prunus (Rosaceae) based on RAD-seq and genome skimming analyses
Na Sua,b,1, Richard G.J. Hodelc,1, Xi Wanga,b,1, Jun-Ru Wanga,b,1, Si-Yu Xiea,b, Chao-Xia Guia,b, Ling Zhangd, Zhao-Yang Changa,b, Liang Zhaoa,b,*, Daniel Pottere, Jun Wenc     
a. College of Life Sciences, Northwest A&F University, Yangling 712100, China;
b. Herbarium of Northwest A&F University, Yangling 712100, China;
c. Department of Botany, National Museum of Natural History, MRC 166, Smithsonian Institution, Washington, DC 20013-7012, USA;
d. College of Life Sciences, Tarim University, Alaer 843300, China;
e. Department of Plant Sciences, MS2, University of California, Davis, CA 95616, USA
Abstract: Prunus is an economically important genus widely distributed in the temperate Northern Hemisphere. Previous studies on the genus using a variety of loci yielded conflicting phylogenetic hypotheses. Here, we generated nuclear reduced representation sequencing data and plastid genomes for 36 Prunus individuals and two outgroups. Both nuclear and plastome data recovered a well-resolved phylogeny. The species were divided into three main clades corresponding to their inflorescence types, - the racemose group, the solitary-flower group and the corymbose group - with the latter two sister to one another. Prunus was inferred to have diversified initially in the Late Cretaceous around 67.32 million years ago. The diversification of the three major clades began between the Paleocene and Miocene, suggesting that paleoclimatic events were an important driving force for Prunus diversification. Ancestral state reconstructions revealed that the most recent common ancestor of Prunus had racemose inflorescences, and the solitary-flower and corymb inflorescence types were derived by reduction of flower number and suppression of the rachis, respectively. We also tested the hybrid origin hypothesis of the racemose group proposed in previous studies. Prunus has undergone extensive hybridization events, although it is difficult to identify conclusively specific instances of hybridization when using SNP data, especially deep in the phylogeny. Our study provides well-resolved nuclear and plastid phylogenies of Prunus, reveals substantial cytonuclear discord at shallow scales, and sheds new light on inflorescence evolution in this economically important lineage.
Keywords: Prunus    Rosaceae    RAD-Seq    Chloroplast genome    Hybridization    Inflorescence evolution    
1. Introduction

Angiosperms exhibit a great diversity of inflorescence types, such as racemes, panicles, corymbs, and cymes, depending on the number and arrangement of flowers on shoots, and their branching patterns (Benlloch et al., 2007; Endress, 2010). The evolution of inflorescences has attracted the interest of botanists for a long time (e.g., Parkin, 1914; Stebbins, 1973; Takhtajan, 1991; Endress, 2010). The rapid development of DNA sequencing technology, computational resources, and phylogenetic inference methods provide an opportunity for inferring accurate phylogenies and conducting rigorous tests of hypotheses concerning inflorescence evolution for some angiosperm groups based on genomic and phylogenetic data (Endress, 2010; Gerrath et al., 2017; Ma et al., 2017).

Prunus L. consists of 250–400 species of trees and shrubs widely distributed in the northern temperate zone and subtropical and tropical regions (Rehder, 1956; Yü et al., 1986; Lu et al., 2003), with eastern Asia being its center of diversity (Wen et al., 2008; Chin et al., 2014). This genus is characterized by simple leaves with stipules and leaf glands, a superior single ovary, and drupe fruits (Lu et al., 2003; Zhao et al., 2016). Many species of Prunus are economically significant as food crops, such as peach, plum, almond, and sweet cherry (Fig. 1; Bortiri et al., 2001; Lee and Wen, 2001), and many others have been used as ornamentals, timber, and medicine (Andro and Riffaud, 1995; Lee and Wen, 2001; Wen et al., 2008). Species of Prunus show a diversity of floral displays with several distinct inflorescence types, including racemes, corymbs, and solitary flowers, making it an ideal group to investigate the evolutionary transitions among different inflorescence organizations.

Fig. 1 Morphological diversity of Prunus species. (A) P. triloba; (B) P. triloba f. multiplex; (C) P. serrulata; (D) P. laurocerasus; (E) P. cerasifera f. atropurpurea; (F) P. mume; (G) P. zippeliana; (H) P. glanulosa; (I) P. serrulata 'Grandiflora'; (J) P. padus; (K) P. serotina; (L) P. henryi; (M–N) P. hypoxantha; (O) P. tomentosa; (P–R) P. maackii; (S) cherry; (T) apricot; (U) plum; (V) peach.

Rehder (1956) divided Prunus into five subgenera, namely Amygdalus L., Cerasus Mill., Lauro-cerasus Tourn. ex Duhamel, Padus Mill. and Prunus, and several taxonomists have accepted this classification (Kalkman, 2004; Wen et al., 2008; Chin et al., 2010, 2014; Zhao et al., 2016). Over the last two decades, researchers have tried to reconstruct the phylogenetic relationships of Prunus using different genomic regions, such as plastid markers, nuclear ribosomal ITS, and other nuclear loci (Bortiri et al., 2001, 2002, 2006; Lee and Wen, 2001; Wen et al., 2008; Chin et al., 2010, 2014; Shi et al., 2013; Zhao et al., 2016, 2018; Hodel et al., 2021). These analyses have recovered three main groups within Prunus, each of them presenting a synapomorphic inflorescence structure: (1) the solitary-flower group consisting of Prunus subg. Amygdalus (L.) Foche and P. subg. Prunus (including section Armeniaca (Scop.) Nakai); (2) the corymbose group, which includes P. subg. Cerasus (Mill.) Pers; and (3) the racemose group comprising P. subg. Lauro-cerasus and P. subg. Padus, as well as the Maddenia Hook. f. & Thomson and Pygeum Gaertn. groups (Chin et al., 2014). Most species of the solitary-flower and corymbose groups are diploid, whereas taxa of the racemose group usually have higher ploidy levels (Chin et al., 2014; Zhao et al., 2016; Hodel et al., 2021; CCDB, http://ccdb.tau.ac.il/). Nevertheless, the frequent, ancient hybridization events in the evolutionary history of Prunus have greatly challenged the clarification of the phylogenetic relationships among these three groups. The monophyly of the racemose group has been supported by plastid sequences (e.g., Chin et al., 2010, 2014); however, most analyses of nuclear sequence data have resolved the racemose group as paraphyletic (e.g., Bortiri et al., 2001, 2002; Lee and Wen, 2001; Chin et al., 2014), leading to hypotheses of multiple hybrid origins of this mostly polyploid lineage (Chin et al., 2014; Zhao et al., 2016). The monophyly of the racemose group has been strongly supported by hundreds of single-copy nuclear genes and chloroplast genomes from 21 transcriptomic data points (Hodel et al., 2021); however, only two species of the racemose group were included in their study, questioning the representativeness of the sampling regarding the taxonomic and morphological diversity in this lineage. Therefore, these results need to be further tested using more comprehensive taxon sampling.

Due to limited taxon sampling and informative loci in previous molecular phylogenetic studies of Prunus, the phylogenetic relationships among and within the major lineages still need to be reevaluated. Although phylogenomic data from transcriptomes and plastomes have generated a well-supported backbone for Rosaceae, limited taxon sampling of Prunus in these analyses failed to untangle all relationships (Xiang et al., 2016; Zhang et al., 2017). Moreover, the divergence times of some lineages within Prunus estimated in previous studies need to be further tested. Thus, exploring phylogenetic relationships within Prunus with extensive taxon sampling and molecular characters is essential.

The plastid genome (plastome) is characterized by its conserved structure and typically maternal inheritance, which makes it helpful in reconstructing phylogenetic relationships at both shallow and deep taxonomic levels (e.g., Li et al., 2019; 2021a; Liu et al., 2019, 2020a; 2020b; Li et al., 2020; Thode et al., 2020; Walker et al., 2022; Wang et al., 2020; Bai et al., 2021; Lei et al., 2021; Li et al., 2021b; Wu et al., 2021; Su et al., 2021; Lee et al., 2022; Liu et al., 2022a, 2022b; Xu et al., 2022; Zhang et al., 2022). However, the uniparental inheritance of plastomes limits their power to fully elucidate the evolutionary histories of lineages with pervasive reticulate evolution, such as the family Rosaceae. Therefore, nuclear and plastome data are needed to infer robust phylogenetic reconstruction. Restriction-site associated DNA sequencing (RAD-seq) can readily generate thousands of unlinked nuclear loci that can resolve dense reticulation on shallow systematic levels (e.g., Vargas et al., 2017; Ma et al., 2018; Mu et al., 2020; Zhou et al., 2020; Hodel et al., 2022).

In this study, we used plastomes from genome skimming data and single nucleotide polymorphism (SNPs) data from RAD-seq data to infer the phylogeny of Prunus, and we also reconstructed the evolution of inflorescence types in Prunus. We aim to (1) resolve phylogenetic relationships among the major lineages, (2) detect potential hybridization events, (3) estimate divergence times, and (4) infer the ancestral states and evolutionary transitions among the states of inflorescence types in Prunus.

2. Materials and methods 2.1. Sampling, DNA extraction, library preparation and sequencing

We sampled 38 accessions, including 32 Prunus species and two outgroup species (Table 1), for specific-locus amplified fragment sequencing (SLAF-seq), which is a reduced-representation genomic sequencing approach. Our sampling was taxonomically and morphologically representative, as these 32 Prunus species represented all subgenera and inflorescence types of Prunus. These species are mainly from Asia, which is considered the origin and diversity center of Prunus (Chin et al., 2014). SLAF-seq is a version of RAD-seq that uses a pre-determined reduced representation scheme to optimize and evenly space loci (Sun et al., 2013). For the plastome data, our sampling was identical to the sampling used for nuclear data, with 37 accessions newly sequenced for this study and one (Prunus davidiana (Carrière) Franch., accession number: MH460864) downloaded from GenBank. Total genomic DNA was extracted from the silica-gel-dried leaves using the CTAB method (Doyle and Doyle, 1987).

Table 1 Voucher information and GenBank accession numbers of sampled species of Prunus and outgroups.
Major group Taxon Voucher Location Inflorescence type Latitude (N) Longitude (E) Altitude (m) SRA accession number
RAD-seq plastome
Armeniaca Prunus mandshurica WX201 Jilin, China simple flower 41°44′49″ 125°59′13″ 377 SRR17479199 SRR12920660
Prunus mume WX206 Yunnan, China simple flower 25°08′40″ 102°44′28″ 1926 SRR17479193 SRR12920640
Prunus sibirica WX205 Jilin, China simple flower 41°43′42″ 125°57′18″ 381 SRR17479209 SRR12920641
Prunus s.str. Prunus ussuriensis WX209 Jilin, China simple flower 41°44′49″ 125°59′13″ 377 SRR17479200 SRR12927898
Prunus salicina SN505 Shaanxi, China simple flower 34°16′20″ 108°05′04″ 355 SRR17479201 SRR17543968
Prunus cerasifera SN506 Shaanxi, China simple flower 34°16′20″ 108°05′04″ 355 SRR17479207 SRR17543970
Amygdalus Prunus davidiana WX207 Shaanxi, China simple flower 34°07′17″ 107°53′46″ 975 SRR17479186 SRR12920639
Prunus davidiana SN501 Shaanxi, China simple flower 34°16′20″ 108°05′04″ 355 SRR17479210
Cerasus Prunus maximowiczii WX215 Jilin, China corymb 41°44′49″ 125°59′13″ 377 SRR17479204 SRR12920657
Prunus discadenia SN502 Shaanxi, China raceme 33°28′60″ 108°29′47″ 2321 SRR17479188 SRR12927899
Prunus serrulata WX211 Jilin, China corymb 41°44′49″ 125°59′13″ 376 SRR17479177 SRR12920636
Prunus tomentosa WX216 Shaanxi, China simple flower 34°16′20″ 108°05′04″ 355 SRR17479198 SRR12920656
Prunus cerasoides WX212 Yunnan, China corymb 25°08′40″ 102°44′28″ 1929 SRR17479203 SRR12927897
Prunus japonica var. nakaii SN503 Jilin, China simple flower 41°43′42″ 125°57′18″ 377 SRR17479211 SRR17543972
Prunus stipulacea WX213 Shaanxi, China corymb 33°23′54″ 108°22′18″ 2320 SRR17479178 SRR12920635
Padus Prunus obtusata WX224 Sichuan, China raceme 30°03′15″ 101°57′47″ 2547 SRR17479197 SRR12920649
Prunus virginiana WX220 Shaanxi, China raceme 34°16′20″ 108°05′04″ 355 SRR17479195 SRR12920652
Prunus serotina WX204 Washington DC, USA raceme 43°54′48″ −77°02′47″ 120 SRR17479181 SRR12920648
Prunus padus WX222 Shaanxi, China raceme 34°07′17″ 107°53′46″ 1251 SRR17479174 SRR17543971
Prunus maackii WX221 Liaoning, China raceme 41°46′31″ 123°25′36″ 26 SRR17479176 SRR12920651
Prunus napaulensis WX225 Sichuan, China raceme 29°31′56″ 103°20′09″ 2543 SRR17479179 SRR12927896
Prunus brachypoda WX223 Hubei, China raceme 31°44′40″ 110°40′33″ 2133 SRR17479208 SRR12920650
Lauro-cerasus Prunus zippeliana WX227 Sichuan, China raceme 29°31′56″ 103°20′09″ 743 SRR17479205 SRR12920646
Prunus laurocerasus WX226 Rockville, Maryland, USA raceme 38°54′48″ −77°00′47″ 119 SRR17479192 SRR12920647
Prunus jenkinsii SN509 Yunnan, China raceme 21°55′11″ 101°16′40″ 570 SRR17479194 SRR17543967
Pygeum Prunus topengii WX229 Guangdong, China raceme 23°03′38″ 113°23′41″ 25 SRR17479190 SRR12920644
Prunus arborea var. montana WX228 Sichuan, China raceme 29°31′56″ 103°20′9″ 745 SRR17479196 SRR12920645
Pygeum macrocarpum SN510 Yunnan, China raceme 23°03′38″ 113°23′41″ 570 SRR17479202 SRR17543969
Maddenia Prunus hypoleuca JR324 Hubei, China raceme 31°25′32″ 110°16′51″ 2134 SRR17479191 SRR13863263
Prunus hypoleuca JR348 Shaanxi, China raceme 34°02′17″ 107°42′12″ 2813 SRR17479180 SRR13863261
Prunus incisoserrata JR354 Gansu, China raceme 34°23′20″ 103°55′44″ 2553 SRR17479189 SRR13863259
Prunus incisoserrata JR440 Shaanxi, China raceme 33°28′59″ 108°29′46″ 2324 SRR17479182 SRR13868097
Prunus wilsonii JR428 Sichuan, China raceme 29°31′46″ 103°20′07″ 2940 SRR17479185 SRR13868095
Prunus wilsonii JR314 Shaanxi, China raceme 33°28′60″ 108°29′46″ 2326 SRR17479184 SRR13868096
Prunus hypoxantha JR426 Sichuan, China raceme 29°31′46″ 103°20′07″ 2950 SRR17479183 SRR13868094
Prunus hypoxantha JR374 Sichuan, China raceme 30°03′40″ 102°00′22″ 2545 SRR17479187 SRR13863257
Outgroups Physocarpus amurensis WX230 Shaanxi, China 34°16′20″ 108°05′04″ 355 SRR17479206 SRR12920643
Prinsepia uniflora WX231 Shaanxi, China 34°16′20″ 108°05′04″ 355 SRR17479175 SRR12920642

For RAD-seq data, the Prunus mume (Siebold) Siebold & Zucc. genome (accession number: GCF_000346735.1) was used in an in silico double enzyme digestion to determine the enzyme pair before preparing the library. The library preparation of RAD-seq was completed with the RsaI + HaeⅢ enzyme pair. The paired-end (264–414 bp) sequencing was carried out on the Illumina Hiseq™ 2500 sequencing platform (Illumina, Inc; San Diego, CA, USA) at Beijing Biomarker Technologies Corporation (Beijing, China). For the plastome data, the libraries were prepared in the Molecular Biology Experiment Center, Germplasm Bank of Wild Species in Southwest China using NEBNext® Ultra™ Ⅱ DNA Library Prep Kit. Paried-end sequencing was conducted on a HiSeq 2500™ (Illumina) in BGI (Shenzhen, China).

2.2. SNP calling and plastome assembly

Raw sequencing reads were assembled de novo using ipyrad v.0.9.74 (Eaton, 2014; Eaton and Overcast, 2020) with default parameters, except the clustering threshold set at 0.85 and the maximum alleles per site set at 4. The resulting assembly consisted of 25, 770, 082 nucleotides per species representing 100, 579 loci. This unfiltered 100, 579 locus alignment was used in subsequent coalescent- and concatenation-based phylogenetic inferences. Additionally, the ipyrad assembly was filtered using vcftools (Danacek et al., 2011) to remove loci with minor allele frequency less than 0.05 (--maf = 0.05), and using several filtering levels for missing data to investigate the impact of missing data on the downstream analyses. We used three filtering levels, allowing a maximum of 40%, 60%, or 80% missing data per site (--max-missing = 0.6, 0.4, 0.2). These three SNP matrices were also used in the downstream phylogenetic analyses. We used BLAST searches to filter the RAD-seq data to remove any chloroplast- and mitochondrion-related loci with references from NCBI GenBank (P. mume chloroplast genome (accession number: NC_023798.1) and mitochondrial genome (accession number: NC_060491.1)).

Raw reads generated by genome skimming (Zhang et al., 2015; Liu et al., 2021) were assembled into plastomes using the GetOrganelle pipeline (Jin et al., 2020). The word sizes were set as 125 bp, round numbers were 15, and k-mers were 75, 85, 95 and 105. Complete plastid genomes from the following species were used as seed sequences for the assembly of all sequenced accessions: Prunus kansuensis Rehder (accession number: NC_023956), Prunus maximowiczii Rupr (accession number: NC_026981), P. mume (accession number: NC_023, 798), Prunus padus L. (accession number: NC_026982), Prunus persica (L.) Batsch (accession number: NC_014697), P. pseudocerasus Lindl (accession number: NC_030599). and P. yedoensis Matsum (accession number: NC_026980). All the plastomes were successfully assembled. The seven published sequences were also used as references for annotating the plastomes with PGA (Qu et al., 2019). The annotated sequences were manually revised in Geneious v.11.0.2.

2.3. Phylogenetic analyses

Phylogenetic relationships were inferred for each of these four data sets (full sequence matrix, and those allowing a maximum of 40%, 60%, or 80% missing data per site). Concatenated Maximum Likelihood (ML) phylogenies were inferred using RAxML v.8.2.12 with 100 bootstrap replicates and 20 independent ML searches. The GTRGAMMA model of evolution was used for the full sequence matrix; the ASC_GTRGAMMA model with the Lewis ascertainment correction was used for the three SNP data sets.

Species trees were constructed using the nuclear data in a coalescent framework with SVDQuartets (Chifman and Kubatko, 2014). All quartets were evaluated for each of the four data matrices, and 100 bootstrap replicates were implemented to assess confidence in phylogenetic relationships. The 100, 579 (~250 bp) loci were considered unlinked genes for the full sequence matrix; however, each SNP was considered unlinked in the three SNP data sets.

All the plastome sequences were aligned using MAFFT (Katoh and Standley, 2013) implemented in Geneious v.11.0.2 (Kearse et al., 2012). The plastome data set was used to infer the phylogeny of Prunus based on the maximum likelihood (ML) method. The analyses were performed by RAxML-HPC Black Box 8.2.12 (Stamatakis, 2014) with 1000 bootstrap replicates and the GTR + G model via the CIPRES Science Gateway website (Miller et al., 2010). For the plastome data set, we also reconstructed the phylogeny of Prunus using MrBayes v.3.2 (Ronquist et al., 2012) based on Bayesian inference (BI) methods. Bayesian inference was performed using ten million generations, the first 25% of trees were discarded as burn-in, and trees were sampled every 1000 generations.

2.4. Hybridization tests

We tested the hypothesized allopolyploid origin of the racemose group using the software package Hybrid Detector (HyDe; Blischak et al., 2018). HyDe is an approach similar to ABBA-BABA tests, and uses phylogenetic invariants arising under a coalescent model to identify hybridization among three ingroup taxa polarized by an outgroup. Within each quartet examined, the parameter γ represents the probability of one ingroup species X being sister to species Y, whereas 1-γ represents the probability of species X being sister to species Z. A significant γ value of 0.5 implies that species X is an F1 hybrid resulting from species Y and Z. We ran three separate analyses using this approach. We ran HyDe analyses exhaustively on all combinations of individual species, and then we combined accessions into three groups based on inflorescence type, i.e., solitary, corymbose and racemose. Additionally, we investigated hybridization within each of these three groups. Physocarpus amurensis (Maxim.) Maxim and Prinsepia uniflora Batalin were used as outgroups for all analyses.

2.5. Divergence time estimation

Divergence times were estimated using the plastome matrix of the 38 individuals in BEAST v.2.5.2 (Bouckaert et al., 2019). Based on the published fossil endocarp Prunus wutuensis from Shandong Province in eastern China (Li et al., 2011), a lognormal prior with an offset of 55.0 million years ago (Mya) and a standard deviation of 1 Mya were used to set the stem age of Prunus subg. Cerasus. A normal prior with offset of 35.0 Mya and a standard deviation of 2 Mya were set for the crown age of the racemose group according to the published amber of a staminate flower of Prunus hirsutipetala D.D. Sokoloff, Remizowa et Nuraliev from northwestern Ukraine (Sokoloff et al., 2018). A nodal calibration was set for the crown age of Prunus (Chin et al., 2014), using a normal prior with an offset of 61.5 Mya and a standard deviation of 3.0. The Markov chain Monte Carlo (MCMC) run was conducted for 100, 000, 000 generations. Trees were sampled every 1000 generations. The log files of BEAST analysis were checked with Tracer 1.5 (Rambaut et al., 2018). Results were considered reliable once the effective sampling size (ESS) for all parameters exceeded 200. We used TreeAnnotator to generate the maximum clade credibility tree with 25% of trees discarded as burn-in.

2.6. Ancestral character state reconstruction

We traced the inflorescence evolution of Prunus using the BI tree inferred from the plastome data set. Three inflorescence types were defined (0) simple flowers, (1) corymbs, and (2) racemes. The detailed inflorescence information was obtained from Flora Republicae Popularis Sinicae (http://www.iplant.cn/frps) and observations in the field listed in Table 1. The 'Stochastic Character Mapping' method implemented in Mesquite v.3.51 (Maddison and Maddison, 2018) was employed to reconstruct ancestral character states.

3. Results 3.1. Characteristics of RAD-seq and plastome data sets

RAD-seq yielded 12.9 Gb raw data across the 38 samples included in this study. The total read number ranged from 1, 436, 523 to 5, 977, 937 among all sequenced individuals. The average Q30 percentage was 93.5% and the average GC content was 42.8% (Table S1). The SNP matrices constructed via de novo assembly using ipyrad consisted of a sequence matrix totaling 25, 770, 082 nucleotide sites per species and a SNP matrix of 1, 587, 718 nucleotides per species. Subsequent filtering of the assembly resulted in data matrices of 9, 637, 65, 176, and 448, 083 sites, respectively (Table S2).

The size of Prunus plastomes ranged from 157, 660 bp (P. davidiana) to 159, 000 bp (P. stipulacea Maxim.) in length. Plastomes of all Prunus species we studied had a quadripartite structure, including a large single-copy (LSC, 85, 764–87, 692 bp) region, a small single-copy (SSC, 18, 855–19, 161 bp) region, and two inverted repeat (IR, 26, 198–26, 458 bp) regions (Table S3). The total GC content of all the Prunus plastomes ranged from 36.6% to 36.8%. All Prunus plastomes encoded 113 unique genes, including 79 protein-coding genes (CDS), four ribosomal RNAs (rRNAs), and 30 transfer RNAs (tRNAs). In addition, 17 genes were duplicated in the IRs, of which six, four, and seven were protein-coding genes, rRNAs and tRNAs, respectively (Table S3). The matrix length of the plastome sequence was 169, 232 bp after trimming.

3.2. Phylogenetic relationships of Prunus

Overall, the nuclear and plastome data both showed three strongly supported major clades (Fig. 2 and Figs. S1–S6), each presenting a characteristic inflorescence architecture: the racemose group (clade 1/A), solitary-flower group (clade 2/B) and corymbose group (clade 3/C). Cytonuclear discord was not detected in the backbone of Prunus. Still, there were some conflicts, especially within the major clades, e.g., for section Armeniaca and section Microcerasus M. Roem. of the solitary-flower group, and within the racemose group, as well as within the corymbose group (Fig. 2). The racemose group (clade 1/A) included the subg. Lauro-cerasus, subg. Padus, and Maddenia and Pygeum groups, and its monophyly was strongly supported with a sister relationship to the clade of the remaining two major groups of Prunus (clades 2/B and 3/C). Padus species formed a monophyletic group based on ipyrad assemblies (Figs. 2, S1, S2 and S4). The nuclear phylogenies inferred using the concatenated-based method were broadly congruent with the phylogenies constructed using a coalescent-based approach (i.e., SVDQuartets) (Figs. 2 and S1–S6). However, the species trees and concatenation-based phylogenies showed a different phylogenetic placement of Prunus serotina Ehrh. Whereas P. serotina was always sister to the remaining Padus species in the SVDQuartets trees (Figs. 2, S1 and S2), this was the case in the concatenation-based phylogeny for only one data set—the one with a maximum of 40% missing data (Fig. S4). In the data sets with no limit on missing data (Figs. S3 and S6), P. serotina was sister to a clade containing all species of Padus and Maddenia. In the data set allowing up to 80% missing data, P. serotina was sister to Maddenia (Fig. S5), which matched the chloroplast topology but not the nuclear coalescent topology.

Fig. 2 The nuclear (left) and plastome (right) topologies for 36 Prunus samples plus two outgroups (Physocarpus amurensis and Prinsepia uniflora). SVDQuartets tree and RAxML tree (left) for 38 species constructed using ipyrad de novo assemblies allowing maximum 60% missing data (65, 176 SNPs). Numbers at nodes indicate bootstrap support values. The support values above the branches show BS (bootstrap support) (left) and PP (posterior probability)/BS (right), and asterisks indicate 1.00/100%. Dashes represent incongruences of BI/ML tree and SVDQuartets/RAxML tree.

In the rest of this section, the tree generated using the ipyrad assembly allowing a maximum of 60% missing data will be used to discuss the phylogenetic relationships of Prunus because there were only minor differences among assemblies (Figs. 2 and S1–S6). Lauro-cerasus was monophyletic in none of the analyses. One species, Prunus laurocerasus L., was sister to all taxa from Maddenia and Padus in the nuclear tree, in contrast with the sister relationship to the Padus species excluding P. serotine, as shown in the plastome tree. Both nuclear and plastome phylogenies showed that Maddenia and Pygeum each formed a monophyletic group (Fig. 2).

The solitary-flower group (clade 2) contained not only species traditionally assigned to it (subg. Amygdalus, section Armeniaca and subg. Prunus s. str.), but also section Microcerasus species of subg. Cerasus. Amygdalus was the first diverging lineage within the solitary flower group, but there were some conflicts between the nuclear and plastome trees within clade 2. Section Microcerasus was a monophyletic clade in the nuclear tree but not in the plastome tree. Armeniaca and Prunus s. str. formed a sister clade, and together they were sister to Microcerasus in the nuclear tree. Armeniaca (Prunus mandshurica (Maxim.) Koehne and P. mume) formed a clade with Microcerasus, and one Armeniaca species (Prunus sibirica L.) was sister to Prunus s. str. in the plastome tree.

The corymbose group was composed of all subg. Cerasus species (clade 3) and was divided into two subclades, but the position of Prunus cerasoides Buch.-Ham. ex D. Don was not congruent in the plastome and nuclear trees. P. cerasoides and P. serrulata Lindl. formed a clade with P. maximowiczii that was siter to the clade of P. stipulacea and P. discadenia Koehne in the nuclear tree. P. cerasoides and (P. stipulacea + P. discadenia) formed a clade that was sister to the clade of P. serrulata and P. maximowiczii in the plastome tree (Fig. 2).

3.3. Hybridization tests

Of the 13, 485 hypotheses of hybridization tested using HyDe, 4180 contained significant evidence of hybridization (Table S4). There was a wide range of γ values among the significant tests, and over 2/3 of γ values were less than 0.3 or greater than 0.7 (Table 2). Typically, γ values that deviate substantially from 0.5, as is the case here, indicate a more ancient history of hybridization within the group, as opposed to a first-generation hybridization event that a γ value of 0.5 would suggest. Species from the racemose group make up the majority of species included in the significant hybridization comparisons, although racemose species are often found included as either parental species or hybrids (Table 2). The HyDe analyses grouping together accessions into three groups suggested that the corymbose group may have originated via hybridization of the racemose and solitary-flower groups (Table S5). HyDe analyses of hybridization within each of the three inflorescence groups indicated high levels of hybridization (i.e., greater than 30%) (Tables S6–S8).

Table 2 The summarized significant HyDe results, representing 4180 hypotheses of hybridization tested out of a total of 13, 485 investigated. The proportion of times that a species from the racemose, solitary, and corymbose group represents Parent 1 (P1), the Hybrid, or Parent 2 (P2) is summarized in the first three rows. The distribution of γ values across all significant tests is shown in the bottom row.
P1 Racemose group (66.4%)
Solitary-flower group (22.4%)
Corymbose group (11.2%)
Hybrid Racemose group (47.5%)
Solitary-flower group (32.0%)
Corymbose group (20.5%)
P2 Racemose group (50.1%)
Solitary-flower group (33.1%)
Corymbose group (16.8%)
γ value γ < 0.3 (34.0%)
0.3 ≤ γ ≤ 0.7 (31.5%)
γ > 0.7 (34.5%)
3.4. Divergence time estimation

The diversification of Prunus began around 67.32 Mya (95% HPD: 55.66–79.77 Mya) in the Late Cretaceous. Clades 2 and 3 split at around 58.38 Mya (95% HPD: 41.23–73.71 Mya), and diversified around 25.92 Mya (95% HPD: 12.83–58.59 Mya) and 11.45 Mya (95% HPD: 3.86–49.58 Mya), respectively (Fig. 3). Clade 1 diversified around 35.41 Mya (95% HPD: 26.51–43.39 Mya). The divergence of Pygeum and Maddenia groups in clade 1 occurred at 12.24 Mya (95% HPD: 2.72–29.29 Mya) and 2.81 Mya (95% HPD: 0.86–12.82 Mya), respectively.

Fig. 3 Chronogram of Prunus based on plastome data sets inferred from BEAST. Blue bars represent the 95% high posterior density credibility interval for node ages. Three calibration points are indicated with red arrows. Nodes of interests were marked as 1–7. Bayesian posterior probabilities are given above branches.
3.5. Ancestral character state reconstruction

Our results indicated that the ancestral inflorescence type of Prunus was the raceme. The ancestral state of clades 2 and 3 in Prunus was a simple flower. As such, the simple flower and corymb states were derived. In the corymbose group, there was an evolutionary reversal to raceme in P. discadenia SN502 (Fig. 4).

Fig. 4 Ancestral character reconstruction of Prunus inflorescences using the Stochastic Character Mapping implemented in the program Mesquite.
4. Discussion

Below, we discuss the main findings of our phylogenetic results and inferences of hybridization events, divergence times, and inflorescence evolution in Prunus.

4.1. Phylogenetic relationships

Our study integrated nuclear and plastome data to infer phylogenetic relationships of Prunus based on species sampling representing all major lineages within the genus. Data sets from both nuclear and plastid genomes supported that Prunus consisted of three main clades—the racemose group, the solitary-flower group, and the corymbose group. The corymbose group was sister to the solitary-flower group, and then together sister to the racemose group with strong support. This topology was consistent with previous findings based on plastid genes (Chin et al., 2010, 2014; Zhao et al., 2016) and those from some single-copy nuclear genes and plastomes (Hodel et al., 2021), but in conflict with results from other analyses of nuclear genes (Chin et al., 2010, 2014). Although our two data sets resulted in a consistently robust backbone, strong phylogenetic conflict was recovered at some shallow-level nodes within each of the three main groups.

The racemose group was previously considered paraphyletic according to phylogenetic trees using a few nuclear gene loci (Chin et al., 2014; Zhao et al., 2016), but it was monophyletic in both nuclear and plastome trees (Fig. 2 and Hodel et al., 2021). The observed monophyly of the racemose group may have resulted from the majority of nuclear loci included in this study tracking the maternal phylogeny of the racemose group. Often, a few outlier genes with outsized influence may skew phylogenetic relationships, so even when relationships appear strongly supported, underlying cytonuclear or gene tree discord may give false confidence in incorrect relationships (Walker et al., 2022). The uncertain phylogenetic placement of the racemose species P. serotina, with one concatenation phylogeny matching the plastid topology as opposed to the nuclear coalescent topology, may indicate that the maternal phylogeny readily dominated some of the SNP data sets when constructing concatenated trees (Fig. S5). In any case, the origin of the racemose group needs to be better explored with single-copy nuclear data that track multiple homologous loci of a specific gene (Zimmer and Wen, 2015) and/or whole nuclear genome data. Both Pygeum and Maddenia groups were strongly supported as monophyletic, which is consistent with previous results (Zhao et al., 2018; Su et al., 2021). The paraphyly of subg. Padus and subg. Lauro-cerasus was suggested in previous studies (Bortiri et al., 2002; Wen et al., 2008; Chin et al., 2010, 2014; Liu et al., 2013; Shi et al., 2013; Zhao et al., 2016, 2018). Our results based on plastome data further confirm these conclusions. Padus and Lauro-cerasus may have each derived from multiple hybridization and allopolyploidy events (Zhao et al., 2016). However, the three trees based on ipyrad de novo assemblies show a monophyletic Padus, and the difference in the position of P. serotina was the most obvious incongruence among the topologies obtained with de novo assembly and those published in previous studies (Liu et al., 2013; Chin et al., 2014; Zhao et al., 2018) in nuclear trees.

Prunus maackii Rupr., a species mainly distributed in northeast China, Korea, and eastern Russia, has been treated as a member of subg. Padus (Rehder, 1956; Yü et al., 1986; Lu et al., 2003). This species has racemose and deciduous inflorescences similar to some species of Padus (Rehder, 1956; Wen et al., 2008). Yet it resembles some members of Cerasus with highly incised stipules (Liu et al., 2013) and a short inflorescence. P. maackii has been supported to be nested within subg. Cerasus based on nuclear and plastid sequences (Lee and Wen, 2001; Bortiri et al., 2006; Wen et al., 2008; Shi et al., 2013; Liu et al., 2013; Chin et al., 2014; Zhao et al., 2016). This species was reported to hybridize with P. maximowiczii of the Cerasus group and formed natural hybrids (Wen et al., 2008; Liu et al., 2013; Shi et al., 2013). Thus, some researchers suggested P. maackii be placed in subg. Cerasus (Wen et al., 2008; Liu et al., 2013). Yet, we found that it formed a clade with Padus species in both nuclear and plastome trees with high support for the first time (Fig. 2), which challenges previous studies. Nevertheless, P. maackii is similar to most members of subg. Padus, and they share two morphological characteristics, i.e., terminal racemes with a few leaves at the base of their peduncle and leaves without glands on the petiole. Our results hence support the placement of P. maackii in subg. Padus.

The solitary-flower group showed several conflicts between plastome and nuclear trees (Fig. 2). Section Armeniaca was monophyletic in the nuclear tree, but not in the plastome tree. Due to maternal inheritance of plastomes and biparental inheritance of nuclear data, section Armeniaca may be derived from one paternal parent and several maternal lineages.

Even though the corymbose group refers to all species of subg. Cerasus, species of Microcerasus were scattered in the solitary-flower group in our analyses. Cerasus, excluding section Microcerasus, was monophyletic and had some cytonuclear discord among Cerasus species (Fig. 2). Microcerasus was regarded as a section of subgenus Cerasus by Rehder (1956) and as a subgenus of the more narrowly defined genus Cerasus by Yü et al. (1986). Microcerasus comprises shrubby and woody species with three axillary buds and a short pedicel, similar to the solitary-flower group members (Lee and Wen, 2001; Wen et al., 2008; Shi et al., 2013). However, true cherries share only one axillary bud at each leaf axil. Moreover, it was reported that Microcerasus species more readily hybridized with species of the solitary-flower group than with those of the Cerasus group (Kataoka et al., 1988).

Two decades ago, Lee and Wen (2001) were the first to systematically investigate the phylogenetic relationships of Prunus based on ITS sequences, and showed that two Microcerasus species had been closely related to the solitary-flower group rather than to the corymbose group. This result was supported by those inferred from different molecular markers in many subsequent studies (Bortiri et al., 2001, 2002, 2006; Wen et al., 2008; Liu et al., 2013; Shi et al., 2013; Chin et al., 2014; Zhao et al., 2016). Our nuclear and plastid results revealed that the Microcerasus species (Prunus tomentosa Thunb. and P. japonica var. nakaii (H. Lév.) Rehder) were nested within the solitary-flower group, which is congruent with previous studies. Our analysis did not support the inclusion of Microcerasus into Cerasus. We found that the ovule developmental characteristics of P. tomentosa were more similar to members of the solitary-flower group than to Cerasus (personal observation by L. Zhao). Shi et al. (2013) also suggested that the Cerasus group only included true cherries, and that Microcerasus did not belong to it. Based on morphological and molecular evidence, it seemed reasonable to assign Microcerasus to the solitary-flower group. Mowrey and Werner (1990) pointed out that due to its paraphyly in all analyses Microcerasus was not a clade, which has been supported by subsequent studies (Lee and Wen, 2001; Bortiri et al., 2002; Wen et al., 2008; Chin et al., 2014). Our results also show that Microcerasus species did not form a clade in the plastome tree. Therefore, further studies should sample additional Microcerasus species and closely related taxa to better understand their origin(s) and systematic placement.

4.2. On the hybrid origins of major Prunus lineages

Zhao et al. (2016) hypothesized that the racemose group derived from multiple hybridization events and allopolyploid speciation events. The maternal parent may have been extinct, and the paternal parents shared common ancestry with various members of the corymbose-solitary flower lineages, leading to the observed conflicts between the plastid and nuclear topologies (Zhao et al., 2016). However, our results support the monophyly of the racemose group (Fig. 2).

Although these results challenge the hypothesis of multiple hybrid origins of the racemose group, the HyDe analysis based on RAD-seq data detected frequent hybridization events in this group. In concert with previous studies, the HyDe results indicate pervasive hybridization and/or allopolyploidy within Prunus, especially within the racemose group (Table 2). Our goal was to test the hypothesized allopolyploid origin of the racemose group, but it is challenging to use hybrid detection methods based on extant taxa to assess ancient hybridization that may have involved extinct lineages. Furthermore, subsequent introgression and/or repeated hybridization can lead to conflicting genomic signals in different species. Nonetheless, a large number of hybridization events detected by HyDe highlights the frequency of hybridization within this genus. Analyses of individual species and within groups indicate that hybridization was pervasive throughout Prunus, and our data do not readily confirm any specific hypotheses of ancient hybridization (Tables S5–S8). Because many species in the racemose group are polyploid, HyDe may pick up signals of genome doubling (i.e., allopolyploidy) instead of homoploid hybridization.

Previous studies with less representation of racemose species (Hodel et al., 2021) found that approximately 1% of hybridization tests were significant, in contrast with the present study in which 31% of the tests were significant. A similar approach in another Rosaceae lineage with frequent hybridization (the apple genus Malus) found that 24% of hybridization tests were significant (Liu et al., 2022a). A study of the Hawaiian genus Myrsine using similar markers as the present study (i.e., RAD-seq) found similarly high levels of hybridization using HyDe (between 27.5% and 30.4% of tests significant, depending on the data set; Appelhans et al., 2020). RAD-seq data have been used successfully for non-polyploid taxa with fewer reticulation events (e.g., Hodel et al., 2022). However, RAD-seq loci cannot be used for accurately estimating ancient hybridization events between the polyploid species and its close diploid relatives (e.g., Wang et al., 2021), especially in the presence of extinction events, genome doubling and/or subsequent recent hybridizations.

Our HyDe analyses with accessions grouped based on inflorescence type showed a potential signal of ancient hybridization of the corymbose group, i.e., between the racemose group and solitary-flower group (Table S5), in contrast to the hypothesis of allopolyploid origins of the racemose group suggested by previous studies (Chin et al., 2014; Zhao et al., 2016). While homoploid hybrid origin of the corymbose group with a maternal 'Solitary flower' lineage and a diploid paternal ancestral 'Racemose' lineage is plausible (Chin et al., 2014), it is not supported by the higher ploidy levels of all extant members of the racemose group, nor is it consistent with patterns of cytonuclear discord concerning the relationships among the three major groups observed in previous studies (Chin et al., 2014; Zhao et al., 2016) or the lack of cytonuclear discord concerning those relationships observed here. Moreover, our HyDe analyses using extant taxa grouped based on inflorescence types could not explicitly test Zhao et al.'s (2016) hypothesis of multiple allopolyploid origins of the racemose group, because that hypothesis invoked members of a now-extinct lineage as the maternal parents in the hybridization events. In fact, if an ancestral member of the corymbose group acted as one or more of paternal parents in the hybridization events, as suggested by Zhao et al. (2016), then one would expect exactly the results seen in our species group HyDe tests here, because some nuclear genes would group the corymbose species with their sister solitary-flower group and others would group them with the hybrid racemose lineages to which they gave rise via hybridization. On the other hand, the corymbose inflorescence seems to represent an intermediate morphological state between the raceme and the simple flowers, which may be consistent with the above hybridization hypothesis of the corymbose group. Future additional studies are needed to rigorously test these competing hypotheses using phylogenomic and developmental morphological data.

4.3. Temporal diversification of main clades in Prunus

A previous study based on four plastid genes and ITS sequences estimated that the ancestor of Prunus emerged between 56.7 and 67.4 Mya (Chin et al., 2014). The age of Prunus in the present study slightly predates their estimate (Fig. 3). In our chronogram, the ancestor of clades 2 and 3 firstly split around 58.38 Mya (95% HPD: 41.23–73.71 Mya). This is congruent with the fossil record of P. wutuensis, which mostly resembled extant P. yedoensis of subgenus Cerasus (Li et al., 2011). The evolution of plants has often been impacted by paleoclimatic and geologic events. During the boundaries between epochs of the Paleogene, key paleoclimatic events caused multiple effects on the evolution and distribution of plants. Around 55.8 Mya, there was an abrupt period of global warming caused by a transient burst of CO2, known as the Paleocene-Eocene Thermal Maximum (PETM) (Currano et al., 2008; McInerney and Wing, 2011). During this period, floristic changes have been shown to occur in response to climate fluctuations (Wing and Currano, 2013). Clades 2 and 3 split around 58.38 Mya, which corresponds to the PETM. Therefore, our results seem to indicate that climatic change may have influenced the diversification of Prunus during the PETM.

In addition, the Eocene–Oligocene transition (EOT, 30–40 Mya) in the Cenozoic witnessed a global cooling and led to a reorganization of organisms (Prothero, 1994; Sun et al., 2014; Deng et al., 2020). The diversification of clades 1 and 2 was herein estimated to be around 35.41 Mya (95% HPD: 26.51–43.39 Mya) in the Late Eocene and 25.92 Mya (95% HPD: 12.83–58.59 Mya) in the Late Oligocene, respectively (Fig. 3). There was another climatic cooling event after the Middle Miocene Climatic Optimum (MMCO, around 15 Mya) (Flower and Kennett, 1994). The molecular dating analysis indicate that clade 3 of Prunus initially diverged when a climatic cooling event occurred after MMCO. The divergence time of the three clades coincided with the two cooling events, which also indicates that the paleoclimatic events might have impacted Prunus diversification and evolution.

4.4. Ancestral character state reconstruction of inflorescences

By tracking the evolution of inflorescence types on the Prunus phylogeny, we inferred that the raceme was the ancestral state of inflorescence in this genus. The solitary flower and corymbose inflorescence types might be derived from the reduction of the flower number and suppression of the rachis, respectively. This result is consistent with the hypothesis that racemes are a primitive state in angiosperms (Stebbins, 1973). Although terminal solitary flowers (Parkin, 1914) and panicles (Takhtajan, 1991) have also been regarded as an ancestral state of inflorescences, there has been some skepticism about the reliability of any of these ideas because the earliest-diverging angiosperms, the ANITA grade, do not possess any panicles, and single flowers terminal on shoots are only present in several genera, such as Austrobaileya C.T. White (in part), Schisandra Michx (in part), and Illicium L. (Endress, 2010).

Simple flowers were the ancestral state of clades 2 and 3 of Prunus based on mature inflorescence types. In fact, a lateral floral primordium was initiated but it did not develop further, contributing to simple flowers at maturity in P. persica (personal observation by L. Zhao). P. salicina and P. americana Marshall have a typical inflorescence type with 2–3 flowers clustered in a bud (personal observation by L. Zhao and J. Wen), which resemble a corymb more than a simple flower. Such developmental/morphological evidence suggests that there may be a transition between simple flowers and corymbs. Hence, we assume the ancestral state of clades 2 and 3 had a transitional inflorescence type, which shows simple flowers only at maturity, with more than one flower observed even then. Corymb and simple flowers in Prunus were formed by continuing and stopping development of lateral floral primordium from their ancestor, respectively. Furthermore, P. discadenia SN502, a member of subg. Cerasus, had racemose inflorescence. It seems that this represents a case of reverse evolution in Prunus, i.e., racemes likely evolved from corymbs by elongation of the rachis. However, we only analyzed the mature character state of Prunus and did not provide robust evidence for developmental evolutionary histories of inflorescences in this study. Future studies integrating inflorescence development and phylogenetic analyses are needed to understand the evolutionary directionality of Prunus inflorescences better.

5. Conclusions

This study sheds light on the phylogenetic relationships, hybridization events and inflorescence evolution of Prunus, an economically important plant group. RAD-seq data have been successfully used to elucidate the phylogenetic relationships of various polyploid lineages, such as bamboos (Guo et al., 2021) and Salix L. (Wagner et al., 2020). However, RAD-seq data cannot distinguish orthologs from paralogs for the polyploid species, limiting their utility for tracing ancient hybridization events. Thus, additional studies using whole-genome sequencing or genome resequencing including more extensive sampling of the racemose group and their allies are needed to test the competing hypotheses of multiple allopolyploid origins of the racemose group vs. the homoploid hybrid origin of the corymbose group (Zhao et al., 2016).

Acknowledgments

We thank Prof. Zhong-Hu Li and Dr. Miao–Miao Ju of Northwest University and Dr. Li Feng of Xi'an Jiaotong University, Dr. Fu-Zhen Guo and Xiao-Hua He from Instrument Sharing Platform of Northwest A&F University and Molecular Biology Experiment Center, Germplasm Bank of Wild Species in Southwest China for help with data analysis, and Mr Hai-Ning Li, Xiao-Jie Li and Prof. You Zhou for providing plant photographs (Fig. 1LO). We also thank Dr. Florian Jabbour and Dr. Valéry Malécot for their valuable comments. Computational analyses were partially conducted on the Smithsonian Institution High Performance Computing Cluster (SI/HPC, "Hydra": https://doi.org/10.25572/SIHPC). This work was supported by grants from National Natural Science Foundation of China (No. 32170381 and 31770200).

Authors contributions

Na Su: Formal analysis, Investigation, Methodology, Software, Visualization, Writing - original draft, Writing - review & editing. Richard G. J. Hodel: Formal analysis, Investigation, Methodology, Software, Visualization, Writing - original draft, Writing - review & editing. Xi Wang: Formal analysis, Investigation, Methodology, Software, Visualization, Writing- original draft. Jun-ru Wang: Formal analysis, Investigation, Methodology, Software, Visualization, Writing - original draft. Si-yu Xie: Methodology, Software, Visualization, Writing - original draft. Chao-xia Gui: Writing - original draft. Ling Zhang: Writing - original draft. Zhao-yang Chang: Writing - original draft. Liang Zhao: Conceptualization, Funding acquisition, Investigation, Methodology, Project administration, Supervision, Validation, Writing - review & editing. Daniel Potter: Conceptualization, Investigation, Methodology, Project administration, Supervision, Validation, Writing - review & editing. Jun Wen: Conceptualization, Funding acquisition, Investigation, Methodology, Project administration, Supervision, Validation, Writing - review & editing.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.pld.2023.03.013.

References
Andro, M.C., Riffaud, J.P., 1995. Pygeum africana extract for the treatment of patients with benign prostatic hyperplasia: a review of 25 years of published experience. Curr. Ther. Res., 56: 796-817. DOI:10.1016/0011-393X(95)85063-5
Appelhans, M.S., Paetzold, C., Wood, K.R., et al., 2020. RADseq resolves the phylogeny of Hawaiian Myrsine (Primulaceae) and provides evidence for hybridization. J. Syst. Evol., 58: 823-840. DOI:10.1111/jse.12668
Bai, H.R., Oyebanji, O., Zhang, R., et al., 2021. Plastid phylogenomic insights into the evolution of subfamily Dialioideae (Leguminosae). Plant Divers., 41: 27-34.
Benlloch, R., Berbel, A., Serrano-Mislata, A., et al., 2007. Floral initiation and inflorescence architecture: a comparative view. Ann. Bot., 100: 659-676. DOI:10.1093/aob/mcm146
Blischak, P.D., Chifman, J., Wolfe, A.D., et al., 2018. HyDe: a Python package for genome-scale hybridization detection. Syst. Biol., 67: 821-829. DOI:10.1093/sysbio/syy023
Bortiri, E., Oh, S.H., Jiang, J.G., et al., 2001. Phylogeny and systematics of Prunus (Rosaceae) as determined by sequences analysis of ITS and the chloroplast trnL-trnF spacer DNA. Syst. Bot., 26: 797-807.
Bortiri, E., Oh, S.H., Gao, F.Y., et al., 2002. The phylogenetic utility of nucleotide sequences of sorbitol 6-phosphate dehydrogenase in Prunus (Rosaceae). Am. J. Bot., 89: 1697-1708. DOI:10.3732/ajb.89.10.1697
Bortiri, E.S., Vanden, B., Potter, D., 2006. Phylogenetic analysis of morphology in Prunus reveals extensive homoplasy. Plant Syst. Evol., 259: 53-71. DOI:10.1007/s00606-006-0427-8
Bouckaert, R., Vaughan, T.G., Barido-Sottani, J., et al., 2019. BEAST 2.5: an advanced software platform for Bayesian evolutionary analysis. PLoS Comput. Biol., 15: e1006650. DOI:10.1371/journal.pcbi.1006650
Chifman, J., Kubatko, L., 2014. Quartet inference from SNP data under the coalescent model. Bioinformatics, 30: 3317-3324. DOI:10.1093/bioinformatics/btu530
Chin, S.W., Wen, J., Johnson, G., et al., 2010. Merging Maddenia with the morphologically diverse Prunus (Rosaceae). Bot. J. Linn. Soc., 163: 236-245. DOI:10.1111/j.1095-8339.2010.01083.x
Chin, S.W., Shaw, J., Haberle, R., et al., 2014. Diversification of almonds, peaches, plums and cherries–molecular systematics and biogeographic history of Prunus (Rosaceae). Mol. Phylogenet. Evol., 76: 34-38. DOI:10.1016/j.ympev.2014.02.024
Currano, E.D., Wilf, P., Wing, S.L., et al., 2008. Sharply increased insect herbivory during the paleocene–eocene thermal maximum. Proc. Natl. Acad. Sci. U.S.A., 105: 1960-1964. DOI:10.1073/pnas.0708646105
Danacek, P., Auton, A., Abecasis, G., et al., 2011. The variant call format and VCFtools. Bioinformatics, 27: 2156-2158. DOI:10.1093/bioinformatics/btr330
Deng, W.Y.D., Su, T., Wappler, T., et al., 2020. Sharp changes in plant diversity and plant-herbivore interactions during the Eocene–Oligocene transition on the southeastern Qinghai-Tibetan Plateau. Global Planet. Change, 194: 103293. DOI:10.1016/j.gloplacha.2020.103293
Doyle, J.J., Doyle, J.L., 1987. A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochem. Bull., 19: 11-15.
Eaton, D.A.R., 2014. PyRAD: assembly of de novo RADseq loci for phylogenetic analyses. Bioinformatics, 30: 1844-1849. DOI:10.1093/bioinformatics/btu121
Eaton, D.A.R., Overcast, I., 2020. Ipyrad: interactive assembly and analysis of RADseq datasets. Bioinformatics, 36: 2592-2594. DOI:10.1093/bioinformatics/btz966
Endress, P.K., 2010. Disentangling confusions in inflorescence morphology: patterns and diversity of reproductive shoot ramification in angiosperms. J. Syst. Evol., 48: 225-239. DOI:10.1111/j.1759-6831.2010.00087.x
Flower, B.P., Kennett, J.P., 1994. The middle Miocene climatic transition: east Antarctic ice sheet development, deep ocean circulation and global carbon cycling. Palaeogeogr. Palaeoclimatol. Palaeoecol., 108: 537-555. DOI:10.1016/0031-0182(94)90251-8
Gerrath, J.M., Posluszny, U., Ickert-Bond, S.M., et al., 2017. Inflorescence morphology and development in the basal rosid lineage Vitales. J. Syst. Evol., 55: 542-558. DOI:10.1111/jse.12261
Guo, C., Ma, P.F., Yang, G.Q., et al., 2021. Parallel ddRAD and genome skimming analyses reveal a radiative and reticulate evolutionary history of the temperate bamboos. Syst. Biol., 70: 756-773. DOI:10.1093/sysbio/syaa076
Hodel, R.G.J., Zimmer, E., Wen, J., 2021. A phylogenomic approach resolves the backbone of Prunus (Rosaceae) and identifies signals of hybridization and allopolyploidy. Mol. Phylogenet. Evol., 160: 107118. DOI:10.1016/j.ympev.2021.107118
Hodel, R.G.J., Massatti, R., Knowles, L., 2022. Hybrid enrichment of adaptive variation revealed by genotype-environment associations in montane sedges. Mol. Ecol., 31: 3722-3737. DOI:10.1111/mec.16502
Jin, J.J., Yu, W.B., Yang, J.B., et al., 2020. GetOrganelle: a fast and versatile toolkit for accurate de novo assembly of organelle genomes. Genome Biol., 21: 241. DOI:10.1186/s13059-020-02154-5
Kalkman, C., 2004. Rosaceae. In: Kubitzki, K. (Ed. ), The Families and Genera of Vascular Plants VI. Springer, Berlin, pp. 343-386.
Kataoka, I., Sugiura, A., Tomana, T., 1988. Interspecific hybridization between Microcerasus and other Prunus spp. J. Jpn. Soc. Hortic. Sci., 56: 398-407. DOI:10.2503/jjshs.56.398
Katoh, K., Standley, D.M., 2013. MAFFT multiple sequence alignment software version 7: improvement in performance and usability. Mol. Biol. Evol., 30: 772-780. DOI:10.1093/molbev/mst010
Kearse, M., Moir, R., Wilson, A., et al., 2012. Geneious basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics, 28: 1647-1649. DOI:10.1093/bioinformatics/bts199
Lee, S., Wen, J., 2001. A phylogenetic analysis of Prunus and the Amygdaloideae (Rosaceae) using ITS sequences of nuclear ribosomal DNA. Am. J. Bot., 88: 150-160. DOI:10.2307/2657135
Lee, S.Y., Xu, K.W., Huang, C.Y., et al., 2022. Molecular phylogenetic analyses based on the complete plastid genomes and nuclear sequences reveal Daphne (Thymelaeaceae) to be non-monophyletic as current circumscription. Plant Divers., 44: 279-289. DOI:10.1016/j.pld.2021.11.001
Lei, F.W., Tong, L., Zhu, Y.X., et al., 2021. Plastid phylogenomics and biogeography of the medicinal plant lineage Hyoscyameae (Solanaceae). Plant Divers., 43: 192-197. DOI:10.1016/j.pld.2021.01.005
Li, Y., Smith, T., Liu, C.J., et al., 2011. Endocarps of Prunus (Rosaceae: prunoideae) from the early Eocene of wutu, Shandong Province, China. Taxon, 60: 555-564. DOI:10.1002/tax.602021
Li, H.T., Yi, T.S., Gao, L.M., et al., 2019. Origin of angiosperms and the puzzle of the Jurassic gap. Nat. Plants, 5: 461-470. DOI:10.1038/s41477-019-0421-0
Li, Q.J., Su, N., Zhang, L., et al., 2020. Chloroplast genomes elucidate diversity, phylogeny, and taxonomy of Pulsatilla (Ranunculaceae). Sci. Rep., 10: 19781. DOI:10.1038/s41598-020-76699-7
Li, H.T., Luo, Y., Gan, L., et al., 2021a. Plastid phylogenomic insights into relationships of all flowering plant families. BMC Biology, 19: 232. DOI:10.1186/s12915-021-01166-2
Li, X.P., Zhao, Y.M., Tu, X.D., et al., 2021b. Comparative analysis of plastomes in Oxalidaceae: phylogenetic relationships and potential molecular markers. Plant Divers., 43: 281-291. DOI:10.1016/j.pld.2021.04.004
Liu, X.L., Wen, J., Nie, Z.L., et al., 2013. Polyphyly of the Padus group of Prunus (Rosaceae) and the evolution of biogeographic disjunctions between eastern Asia and eastern North America. J. Plant Res., 126: 351-361. DOI:10.1007/s10265-012-0535-1
Liu, B.B., Hong, D.Y., Zhou, S.L., et al., 2019. Phylogenomic analyses of the Photinia complex support the recognition of a new genus Phippsiomeles and the resurrection of a redefined Stranvaesia in Maleae (Rosaceae). J. Syst. Evol., 57: 678-694. DOI:10.1111/jse.12542
Liu, B.B., Campbell, C.S., Hong, D.Y., et al., 2020a. Phylogenetic relationships and chloroplast capture in the Amelanchier-Malacomeles-Peraphyllum clade (Maleae, Rosaceae): evidence from chloroplast genome and nuclear ribosomal DNA data using genome skimming. Mol. Phylogenet. Evol., 147: 106784. DOI:10.1016/j.ympev.2020.106784
Liu, B.B., Liu, G.N., Hong, D.Y., et al., 2020b. Eriobotrya belongs to Rhaphiolepis (Maleae, Rosaceae): evidence from chloroplast genome and nuclear ribosomal DNA data. Front. Plant Sci., 10: 1731. DOI:10.3389/fpls.2019.01731
Liu, B.B., Ma, Z.Y., Ren, C., et al., 2021. Capturing single-copy nuclear genes, organellar genomes, and nuclear ribosomal DNA from deep genome skimming data for plant phylogenetics: a case study in Vitaceae. J. Syst. Evol., 59: 1124-1138. DOI:10.1111/jse.12806
Liu, B.B., Ren, C., Kwak, M., et al., 2022a. Phylogenomic conflict analyses in the apple genus Malus s.l. reveal widespread hybridization and allopolyploidy driving diversification, with insights into the complex biogeographic history in the Northern Hemisphere. J. Integr. Plant Biol., 64: 1020-1043. DOI:10.1111/jipb.13246
Liu, C., Chen, H.H., Tang, L.Z., et al., 2022b. Plastid genome evolution of a monophyletic group in the subtribe Lauriineae (Laureae, Lauraceae). Plant Divers., 44: 377-388. DOI:10.1016/j.pld.2021.11.009
Lu, L.D., Gu, C.Z., Li, C.L., et al., 2003. Rosaceae. In: Wu, Z.Y., Raven, P.H., Hong, D.Y. (Eds. ), Flora of China. Beijing, Science Press. Missouri Botanical Garden Press, St. Louis, MO, pp. 46-434.
Ma, Q., Zhang, W.H., Xiang, Q.Y., 2017. Evolution and developmental genetics of floral display-a review of progress. J. Syst. Evol., 55: 487-515. DOI:10.1111/jse.12259
Ma, Z.Y., Wen, J., Tian, J.P., et al., 2018. Testing reticulate evolution of four Vitis species from East Asia using restriction-site associated DNA sequencing. J. Syst. Evol., 56: 311-399. DOI:10.1364/oe.26.000311
Maddison, W.P., Maddison, D.R., 2018. Mesquite: a Modular System for Evolutionary Analysis v3.51 (Version 3.51). http://www.mesquiteproject.org.
McInerney, F.A., Wing, S.L., 2011. The Paleocene–Eocene Thermal Maximum: a perturbation of the carbon cycle, climate, and biosphere with implications for the future. Annu. Rev. Earth Planet Sci., 39: 489-516. DOI:10.1146/annurev-earth-040610-133431
Miller, M.A., Pfeiffer, W., Schwartz, T., 2010. Creating the CIPRES Science Gateway for inference of large phylogenetic trees. In: 2010 Gateway Computing Environments Workshop (GCE). New Orleans, pp. 1-8.
Mowrey, B.D., Werner, D.J., 1990. Phylogenetic relationships among species of Prunus as inferred by isozyme markers. Theor. Appl. Genet., 80: 129-133. DOI:10.1007/bf00224026
Mu, X.Y., Tong, L., Sun, M., et al., 2020. Phylogeny and divergence time estimation of the walnut family (Juglandaceae) based on nuclear RAD-Seq and chloroplast genome data. Mol. Phylogenet. Evol., 147: 106802. DOI:10.1016/j.ympev.2020.106802
Parkin, J., 1914. The evolution of the inflorescence. Bot. J. Linn. Soc., 42: 511-563. DOI:10.1111/j.1095-8339.1914.tb00888.x
Prothero, D.R., 1994. The late Eocene-Oligocene extinctions. Annu. Rev. Earth Planet Sci., 22: 145-165. DOI:10.1146/annurev.ea.22.050194.001045
Qu, X.J., Moore, M.J., Li, D.Z., et al., 2019. PGA: a software package for rapid, accurate, and flexible batch annotation of plastomes. Plant Methods, 15: 1-12. DOI:10.1117/1.jrs.13.048507
Rambaut, A., Drummond, A.J., Xie, D., et al., 2018. Posterior summarization in Bayesian phylogenetics using Tracer 1.7. Syst. Biol., 67: 901-904. DOI:10.1093/sysbio/syy032
Rehder, A., 1956. Manual of Cultivated Trees and Shrubs Hardy in North America Exclusive of the Subtropical and Warmer Temperate Regions, second ed. Macmillan, New York.
Ronquist, F., Teslenko, M., van der Mark, P., et al., 2012. MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol., 61: 539-542. DOI:10.1093/sysbio/sys029
Shi, S., Li, J.L., Sun, J.H., et al., 2013. Phylogeny and classification of Prunus sensu lato (Rosaceae). J. Integr. Plant Biol., 55: 1069-1079. DOI:10.1111/jipb.12095
Sokoloff, D.D., Ignatov, M.S., Remizowa, M.V., et al., 2018. Staminate flower of Prunus s.l. (Rosaceae) from Eocene rovno amber (Ukraine). J. Plant Res., 131: 925-943. DOI:10.1007/s10265-018-1057-2
Stamatakis, A., 2014. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics, 30: 1312-1313. DOI:10.1093/bioinformatics/btu033
Stebbins, G.L., 1973. Evolutionary trends in the inflorescence of angiosperms. Flora, 162: 501-528. DOI:10.1016/S0367-2530(17)31733-4
Su, N., Liu, B.B., Wang, J.R., et al., 2021. On the species delimitation of the Maddenia group of Prunus (Rosaceae): evidence from plastome and nuclear sequences and morphology. Front. Plant Sci., 12: 743643. DOI:10.3389/fpls.2021.743643
Sun, X.W., Liu, D.Y., Zhang, X.F., et al., 2013. SLAF-seq: an efficient method of large-scale de novo SNP discovery and genotyping using high-throughput sequencing. PLoS One, 8: e58700. DOI:10.1371/journal.pone.0058700
Sun, J.M., Ni, X.J., Bi, S.D., et al., 2014. Synchronous turnover of flora, fauna, and climate at the Eocene–Oligocene boundary in Asia. Sci. Rep., 4: 7463. DOI:10.1038/srep07463
Takhtajan, A., 1991. Evolutionary Trends in Flowering Plants. Columbia University Press, New York.
Thode, V.A., Lohmann, L.G., Sanmartín, I., 2020. Evaluating character partitioning and molecular models in plastid phylogenomics at low taxonomic levels: a case study using Amphilophium (Bignonieae, Bignoniaceae). J. Syst. Evol., 58: 1071-1089. DOI:10.1111/jse.12579
Vargas, O.M., Ortiz, E.M., Simpson, B.B., 2017. Conflicting phylogenomic signals reveal a pattern of reticulate evolution in a recent high-Andean diversification (Asteraceae: Astereae: Diplostephium). New Phytol., 214: 1736-1750. DOI:10.1111/nph.14530
Wagner, N.D., He, L., Hörandl, E., 2020. Phylogenomic relationships and evolution of polyploid Salix species revealed by RAD sequencing data. Front. Plant Sci., 11: 1077. DOI:10.3389/fpls.2020.01077
Walker, J.F., Smith, S.A., Hodel, R.G.J., et al., 2022. Concordance-based approaches for the inference of relationships and molecular rates with phylogenomic data sets. Syst. Biol., 71: 943-958. DOI:10.1093/sysbio/syab052
Wang, Y.B., Liu, B.B., Nie, Z.L., et al., 2020. Major clades and a revised classification of Magnolia and Magnoliaceae based on whole plastid genome sequences via genome skimming. J. Syst. Evol., 58: 673-695. DOI:10.1111/jse.12588
Wang, N., Kelly, L.J., McAllister, H.A., et al., 2021. Resolving phylogeny and polyploid parentage using genus-wide genome-wide sequence data from birch trees. Mol. Phylogenet. Evol., 160: 107126. DOI:10.1016/j.ympev.2021.107126
Wen, J., Berggren, S.T., Lee, C.H., et al., 2008. Phylogenetic inferences in Prunus (Rosaceae) using chloroplast ndhF and nuclear ribosomal ITS sequences. J. Syst. Evol., 46: 322-332.
Wing, S.T., Currano, E.D., 2013. Plant response to a global greenhouse event 56 million years ago. Am. J. Bot., 100: 1234-1254. DOI:10.3732/ajb.1200554
Wu, H., Ma, P.F., Li, H.T., et al., 2021. Comparative plastomic analysis and insights into the phylogeny of Salvia (Lamiaceae). Plant Divers., 43: 15-26. DOI:10.1016/j.pld.2020.07.004
Xiang, Y.Z., Huang, C.H., Hu, Y., et al., 2016. Well-resolved Rosaceae nuclear phylogeny facilitates geological time and genome duplication analyses and ancestral fruit character reconstruction. Mol. Biol. Evol., 34: 262-281.
Xu, Y.L., Shen, H.H., Du, X.Y., et al., 2022. Plastome characteristics and species identification of Chinese medicinal wintergreens (Gaultheria, Ericaceae). Plant Divers., 44: 519-529. DOI:10.1016/j.pld.2022.06.002
Yü, T.T., Lu, L.T., Ku, T.C., et al., 1986. Rosaceae (3) Prunoideae. In: Yü, T.T. (Ed. ), Flora Reipublicae Popularis Sinicae, vol. 38. Science Press, Beijing, pp. 1-133.
Zhang, N., Wen, J., Zimmer, E.A., 2015. Congruent deep relationships in the grape family (Vitaceae) based on sequences of chloroplast genomes and mitochondrial genes via genome skimming. PLoS One, 10: e0144701. DOI:10.1371/journal.pone.0144701
Zhang, S.D., Jin, J.J., Chen, S.Y., et al., 2017. Diversification of Rosaceae since the late cretaceous based on plastid phylogenomics. New Phytol., 214: 1355-1367. DOI:10.1111/nph.14461
Zhang, Y.M., Han, L.J., Yang, C.W., et al., 2022. Comparative chloroplast genome analysis of medicinally important Veratrum (Melanthiaceae) in China: insights into genomic characterization and phylogenetic relationships. Plant Divers., 44: 70-82. DOI:10.1016/j.pld.2021.05.004
Zhao, L., Jiang, X.W., Zuo, Y.J., et al., 2016. Multiple events of allopolyploidy in the evolution of the racemose lineages in Prunus (Rosaceae) based on integrated evidence from nuclear and plastid data. PLoS One, 11: e0157123. DOI:10.1371/journal.pone.0157123
Zhao, L., Potter, D., Xu, Y., et al., 2018. Phylogeny and spatio-temporal diversification of Prunus subgenus Laurocerasus section Mesopygeum (Rosaceae) in the Malesian region. J. Syst. Evol., 56: 637-651. DOI:10.1111/jse.12467
Zhou, M.M., Yang, G.Q., Sun, G.L., et al., 2020. Resolving complicated relationships of the Panax bipinnatifidus complex in southwestern China by RAD-seq data. Mol. Phylogenet. Evol., 149: 106851. DOI:10.1016/j.ympev.2020.106851
Zimmer, E.A., Wen, J., 2015. Using nuclear gene data for plant phylogenetics: progress and prospects Ⅱ. Next-gen approaches. J. Syst. Evol., 53: 371-379. DOI:10.1111/jse.12174