Abstract

Background and Aims

Cephalotaxus is a paleo-endemic genus in East Asia that consists of about 7–9 conifer species. Despite its great economic and ecological importance, the relationships between Cephalotaxus and related genera, as well as the interspecific relationships within Cephalotaxus, have long been controversial, resulting in contrasting taxonomic proposals in delimitation of Cephalotaxaceae and Taxaceae. Based on plastome data, this study aims to reconstruct a robust phylogeny to infer the systematic placement and the evolutionary history of Cephalotaxus.

Methods

A total of 11 plastomes, representing all species currently recognized in Cephalotaxus and two Torreya species, were sequenced and assembled. Combining these with previously published plastomes, we reconstructed a phylogeny of Cephalotaxaceae and Taxaceae with nearly full taxonomic sampling. Under a phylogenetic framework and molecular dating, the diversification history of Cephalotaxus and allied genera was explored.

Key Results

Phylogenetic analyses of 81 plastid protein-coding genes recovered robust relationships between Cephalotaxus and related genera, as well as providing a well-supported resolution of interspecific relationships within Cephalotaxus, Taxus, Torreya and Amentotaxus. Divergence time estimation indicated that most extant species of these genera are relatively young, although fossil and other molecular evidence consistently show that these genera are ancient plant lineages.

Conclusions

Our results justify the taxonomic proposal that recognizes Cephalotaxaceae as a monotypic family, and contribute to a clear-cut delineation between Cephalotaxaceae and Taxaceae. Given that extant species of Cephalotaxus are derived from recent divergence events associated with the establishment of monsoonal climates in East Asia and Pleistocene climatic fluctuations, they are not evolutionary relics.

INTRODUCTION

The economically important genus Cephalotaxus Siebold & Zuccarini ex Endlicher consists of about 7–9 evergreen coniferous species occurring in Burma, China, India, Japan, Korea, Laos, Malaysia, Thailand and Vietnam (Fu, 1984; Fu et al., 1999; Farjon, 2010; Lang et al., 2013; Yang, 2017; J. W. Zhang et al., 2019). According to fossil records, the earliest Cephalotaxus species was present in Europe by the middle Eocene (Manchester et al., 2009) and expanded its distribution range into East Asia (north-eastern China) by the late Eocene (He and Tao, 1997). During the Miocene and Pliocene, Cephalotaxus was widespread across the Northern Hemisphere in Eurasia and North America (Miki, 1958; Givulescu, 1973; Mai and Walther, 1978; Wilde 1989; He and Tao, 1997; Kvaček and Walther, 1998; Meller, 1998; Kvaček and Rember, 2000; Walther and Kvaček, 2007). Therefore, the extant members of the genus have a relatively restricted distribution possibly resulting from range contraction and southward migration driven by decreasing global temperatures of the Neogene and Pleistocene glaciation events (Axelrod, 1959; Qian and Ricklefs, 2000; Manchester et al., 2009). Because of its antiquity, Cephalotaxus is an excellent representative of relict plant lineages in East Asia (Manchester et al., 2009), and extant species within this genus are hypothesized to be evolutionary relicts (Fu, 1984; Wu et al., 2005).

Cephalotaxus is distinctive in having two-ovulate bracts in the seed cones (Fu et al., 1999); vegetative characters including general folia morphology and growth habit closely resemble those of Amentotaxus Pilger, Austrotaxus Compton, Pseudotaxus W. C. Cheng, Taxus Linn. and Torreya Arnott. Although the monophyly of Cephalotaxus has been robustly supported by both morphological (Ghimire and Heo, 2014) and molecular evidence (Cheng et al., 2000; Hao et al., 2008; Gao et al., 2015), its taxonomic affinities remain largely disputed. Historically, Cephalotaxus was placed into either Taxaceae (Eichler, 1889; Van Tieghem, 1891; Pilger, 1903; Takhtajan, 1953; Hart, 1987; Price, 1990; Stefanoviac et al, 1998; Christenhusz, 2011) or Cephalotaxaceae (Neger, 1907; Florin, 1948, 1954; Singh, 1961). Due to contradicting taxonomic proposals inferred from different evidence, it remains unresolved whether Cephalotaxaceae should be recognized as a separate family or merged back within Taxaceae. For instance, cladistic analysis based on morphological characters (Ghimire and Heo, 2014) as well as phylogenetic inference utilizing 18S ribosomal DNA sequences (Chaw et al., 1997) and plastid rbcL and matK regions (Quinn et al., 2002) suggest that Cephalotaxaceae should be merged into Taxaceae. However, phylogenetic analyses using a combination of plastid matK and nuclear ribosomal internal transcribed spacer (ITS) regions (Cheng et al., 2000), plastid matK, rbcL, trnL–trnF, psbA–trnH and nuclear ITS (Hao et al., 2008), nuclear LFY and NLY (Lu et al., 2014), and transcriptomic data (Majeed et al., 2018; Ran et al., 2018) support the sister relationship between Cephalotaxaceae and Taxaceae. Nevertheless, a clear-cut delineation between the two families remains unresolved due to the ambiguous position of Amentotaxus and Torreya. Although the analyses by Ran et al. (2018) and X. Zhang et al. (2019) showed that these two genera are closer to Taxaceae, Cheng et al. (2000), Hao et al. (2008), Lu et al. (2014) and Majeed et al. (2018) proposed to include them within Cephalotaxaceae. These conflicts necessitate that the relationships between Cephalotaxus and its relatives require further investigation.

Additionally, the traditional classification of Cephalotaxus is constructed primarily based on folia morphologies, including leaf size, shape of leaves and leaf base, density and arrangement of leaves, etc. (Fu, 1984; Fu et al., 1999; Farjon, 2010; Lang et al., 2013; Yang, 2017; J. W. Zhang et al., 2019). These characters are highly divergent, and are usually overlapping between species (Tripp, 1995), making morphology-based taxonomy particularly perplexing and challenging (J. W. Zhang et al., 2019). Since the establishment of the genus by Endlicher (1847), Cephalotaxus has been the subject of several taxonomic revisions (Fu et al., 1999; Farjon, 2010; Lang et al., 2013; J. W. Zhang et al., 2019). Yet, credible species delineation and unambiguous interspecific relationships within the genus remain unresolved. Given their great economic and ecological importance (Fu, 1984; Lang et al., 2013), resolving the aforementioned problems will help the utilization and conservation of extant Cephalotaxus species, and deepen our understanding of the evolutionary history of gymnosperms.

Phylogenetic reconstruction based on limited sequence variation often results in poor resolution and low support in relationships, particularly at lower taxonomic levels (Rokas and Carroll, 2005; Whitfield and Lockhart, 2007; Philippe et al., 2011). In contrast to Sanger sequencing, next-generation sequencing techniques, which are capable of generating orders of magnitude more data, have offered new approaches to resolving historical problems in plant phylogenetics (e.g. Parks et al., 2009; Barrett et al., 2014; Stull et al., 2015; Zhang et al., 2017; Carlsen et al., 2018; Dong et al., 2018; Li et al., 2019). Plastid genome (plastome) DNA sequences, which harbour a large number of evolutionarily informative variation suitable for phylogenetic analysis, have been widely employed to reconstruct robust evolutionary relationships in phylogenetically challenging plant taxa (e.g. Simpson et al., 2017; Uribe-Convers et al., 2017; Heckenhauer et al., 2019; Ji et al., 2019a).

Here, we sequenced and assembled whole plastomes of nine species of Cephalotaxus and two species of Torreya using a low coverage genome sequencing approach. Based on plastome phylogenomic analyses and fossil-calibrated molecular dating, we aim to (1) elucidate the relationships between Cephalotaxus and related genera; (2) investigate interspecific relationships within Cephalotaxus; and (3) infer the history of species diversification for Cephalotaxus.

MATERIALS AND METHODS

Plant materials, shotgun sequencing, plastome assembly and comparison

According to the most recent taxonomic revisions on Cephalotaxus (Lang et al., 2013; J. W. Zhang et al., 2019), our taxon sampling covered all seven species (Cephalotaxus alpina, C. fortunei, C. griffithii, C. hainanensis, C. harringtonii, C. nana and C. oliveri) recognized in the genus. We also sampled C. mannii and C. sinensis whose taxonomic status as a recognized species differ according to different authors (Fu et al., 1999; Farjon, 2010; Lang et al., 2013; J. W. Zhang et al., 2019), to investigate their relationships to their congeneric species. Voucher specimens (Table 1) were deposited at the herbarium of Kunming institute of Botany, Chinese Academy of Sciences (KUN).

Table 1.

Taxa newly sequenced in this study with source, voucher and GenBank accession numbers

TaxaLocalityVoucherGenBank accession
Cephalotaxus alpinaZixishan Mountain, Chuxiong, Yunnan, ChinaY. Ji and C. Tao 012MT555079
C. fortuneiJinfoshan Mountain, Nanchuan, Chongqing, ChinaY. Ji and C. Liu 084MT555080
C. griffithiiDulongjiang, Gongshan, Yunnan, ChinaGLGS Exp. 38714MT555081
C. hainanensisJianfengling, Ledong, Hainan, ChinaT. Su s. n.MT555082
C. harringtoniiCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 002MT555083
C. manniiCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 003MT555084
C. nanaFengxi, Zhuxi, Hubei, ChinaY. Ji and C. Liu 095MT555085
C. oliveriCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 001MT555086
C. sinensisDengchigou, Baoxing, Sichuan, ChinaY. Ji, W. Chen and T. Su 180MT555087
Torreya fargesii var. yunnanensisCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 005MT555088
T. jackiiCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 004MT555089
TaxaLocalityVoucherGenBank accession
Cephalotaxus alpinaZixishan Mountain, Chuxiong, Yunnan, ChinaY. Ji and C. Tao 012MT555079
C. fortuneiJinfoshan Mountain, Nanchuan, Chongqing, ChinaY. Ji and C. Liu 084MT555080
C. griffithiiDulongjiang, Gongshan, Yunnan, ChinaGLGS Exp. 38714MT555081
C. hainanensisJianfengling, Ledong, Hainan, ChinaT. Su s. n.MT555082
C. harringtoniiCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 002MT555083
C. manniiCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 003MT555084
C. nanaFengxi, Zhuxi, Hubei, ChinaY. Ji and C. Liu 095MT555085
C. oliveriCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 001MT555086
C. sinensisDengchigou, Baoxing, Sichuan, ChinaY. Ji, W. Chen and T. Su 180MT555087
Torreya fargesii var. yunnanensisCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 005MT555088
T. jackiiCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 004MT555089
Table 1.

Taxa newly sequenced in this study with source, voucher and GenBank accession numbers

TaxaLocalityVoucherGenBank accession
Cephalotaxus alpinaZixishan Mountain, Chuxiong, Yunnan, ChinaY. Ji and C. Tao 012MT555079
C. fortuneiJinfoshan Mountain, Nanchuan, Chongqing, ChinaY. Ji and C. Liu 084MT555080
C. griffithiiDulongjiang, Gongshan, Yunnan, ChinaGLGS Exp. 38714MT555081
C. hainanensisJianfengling, Ledong, Hainan, ChinaT. Su s. n.MT555082
C. harringtoniiCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 002MT555083
C. manniiCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 003MT555084
C. nanaFengxi, Zhuxi, Hubei, ChinaY. Ji and C. Liu 095MT555085
C. oliveriCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 001MT555086
C. sinensisDengchigou, Baoxing, Sichuan, ChinaY. Ji, W. Chen and T. Su 180MT555087
Torreya fargesii var. yunnanensisCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 005MT555088
T. jackiiCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 004MT555089
TaxaLocalityVoucherGenBank accession
Cephalotaxus alpinaZixishan Mountain, Chuxiong, Yunnan, ChinaY. Ji and C. Tao 012MT555079
C. fortuneiJinfoshan Mountain, Nanchuan, Chongqing, ChinaY. Ji and C. Liu 084MT555080
C. griffithiiDulongjiang, Gongshan, Yunnan, ChinaGLGS Exp. 38714MT555081
C. hainanensisJianfengling, Ledong, Hainan, ChinaT. Su s. n.MT555082
C. harringtoniiCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 002MT555083
C. manniiCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 003MT555084
C. nanaFengxi, Zhuxi, Hubei, ChinaY. Ji and C. Liu 095MT555085
C. oliveriCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 001MT555086
C. sinensisDengchigou, Baoxing, Sichuan, ChinaY. Ji, W. Chen and T. Su 180MT555087
Torreya fargesii var. yunnanensisCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 005MT555088
T. jackiiCultivated in the Botanical Garden of Kunming Institute of BotanyY. Ji and X. Gong 004MT555089

Total genomic DNA for each species was isolated from approx. 20 mg of silica gel-dried leaf tissues using the CTAB (cetyltrimethylammonium bromide) method (Doyle and Doyle, 1987). Purified genomic DNA (approx. 5 µg) was used to construct shotgun libraries with a TruSeq DNA Sample Prep Kit (Illumina, Inc., USA) following the manufacturer’s instructions. Paired-end sequencing was conducted on the Illumina HiSeq 2500 platform. Shotgun reads were subjected to the NGS QC Toolkit (Patel and Jain, 2012) to remove adaptors and low-quality reads with default parameters. Based on the remaining high-quality Illumina reads, de novo plastome assembly was performed using NOVOPlasty v2.7.0 (Dierckxsens et al., 2017) with a k-mer of 31, and using the large subunit of the Rubisco gene (rbcL) of C. oliveri (AF456387) as the seed for iterative extension of contigs to recover the complete plastome of each species.

The newly generated plastomes were annotated with the Dual Organellar Genome Annotator database (Wyman et al., 2004); the annotation of protein-coding genes was confirmed with a BLAST search against the NCBI protein database. Genes putatively annotated as tRNA were further verified by tRNAscan-SE 1.21 (Schattner et al., 2005) with default parameters. To identify plastome arrangement events in Cephalotaxus, the newly generated plastomes were progressively aligned with those of four species representing closely related genera, Amentotaxus, Torreya, Pseudotaxus and Taxus, using the multiple genome alignment program Mauve 2.3.1 (Darling et al., 2010). Moreover, all Cephalotaxus plastomes were aligned pairwise using the program mVISTA (Mayor et al., 2000) under the LAGAN mode to investigate sequence divergence.

Phylogenetic analyses

In addition to the nine Cephalotaxus plastomes newly generated in this study, a previously sequenced plastome of C. wilsoniana, which was treated as a conspecific variety (Fu et al., 1999; Farjon, 2010) or a synonym of C. harringtonii (Lang et al., 2013; J. W. Zhang et al., 2019), was included in phylogenetic analyses. Moreover, six Cupressaceae plastomes, two Amentotaxus plastomes, eight Torreya plastomes, one Pseudotaxus plastome and 14 Taxus plastomes available in GenBank were incorporated into the dataset (Supplementary data Table S1). According to the backbone relationships of gymnosperms (Christenhusz et al., 2011; Ran et al., 2018), Agathis danunara and Podocarpus totara were selected as outgroups. Eighty-one plastid protein-coding genes commonly shared (Supplementary data Table S2) were aligned with MAFFT v7.4 (Katoh et al., 2013). The best-fitting partition scheme was selected using PartitionFinder v2.1.1 (Lanfear et al., 2017) with the ‘greedy’ search algorithm.

Both maximum likelihood (ML) and Bayesian inference (BI) approaches were used for phylogenetic inference. The ML analyses were conducted with RAxML-HPC BlackBox v8.1.24 (Stamatakis, 2006) with the most suitable model (GTR + G + I) for sequence substitution selected using Modeltest v3.7 (Posada and Crandall, 1998) with the Akaike information criterion (Posada and Buckley, 2004). Ten independent ML searches were performed with 1000 standard bootstrapping replicates. The BI analyses were performed using MRBAYES v 3.1.2 (Ronquist and Huelsenbeck, 2003). Markov chain Monte Carlo (MCMC) runs were initiated with a random tree for 1 million generations, with trees sampled every 100 generations. Trees from the first 250 000 generations were discarded as burn-in. The posterior probability (PP) values were computed based on the remaining trees.

Divergence time estimation

The program BEAST v. 2.4.7 (Bouckaert et al., 2014) was employed to estimate divergence time. We incorporated three fossils to calibrate the molecular clock. The crown age of the lineage including Cupressaceae, Cephalotaxaceae and Taxaceae was set to 197 million years ago (Mya) (Florin, 1958, 1963). The most recent common ancestor of Cephalotaxaceae and Taxaceae was set to a minimum age of 163.5 Mya (Florin, 1963). The stem age of Taxus was set to a minimum age of 100 Mya (Xu et al., 2015). We used the uncorrelated log-normal relaxed molecular clock approach, which allows uncertainty in the age of calibrations to be represented as prior distributions rather than as strict calibration/fixed points. The BEAST analyses were conducted under the sequence substitution model (GTR + G + I) selected by Modeltest with a Yule tree prior. Two independent replications of MCMC simulations were run, sampling every 1000 generations. The stationarity of the chains and convergence of MCMC simulations were monitored by Tracer v1.7.1 (Rambaut et al., 2018). After stationarity was obtained, a maximum credibility tree was constructed with TreeAnotator v1.10 (http://beast.bio.ed.ac.uk/TreeAnnotator), with the first 5000 trees discarded as burn-in.

RESULTS

Features of Cephalotaxus plastomes

Shallow-depth genome sequencing produced approx. 30.1–31.5 million paired-end clean reads per sample. Of those, 4.16 × 105–1.00 × 106 were mapped to the assembled plastomes, with the average sequencing coverage ranging from 464× to 1117× (Supplementary data Table S3). The de novo assemblies generated nine circular Cephalotaxus plastomes, with the sizes of the complete plastome, coding regions and non-coding regions varying from 134 550 to 136 545 bp, 82 928 to 83 985 bp and 51 622 to 52 707 bp, respectively (Supplementary data Table S4). All Cephalotaxus species possessed a similar GC content not only in the whole plastomes assembly (35.1–35.2 %), but also in coding (37.3–37.5 %) and non-coding (31.%–31.6 %) regions (Supplementary data Table S4). All Cephalotaxus plastomes identically encoded 113 unique genes, comprising four rRNA genes, 27 tRNA genes and 82 protein-coding genes (Supplementary data Table S5). In addition, due to the loss of the typical inverted repeat region, it is difficult to define the boundary of large single-copy and small single-copy regions in Cephalotaxus plastomes (Fig. 1).

Map of Cephalotaxus plastomes. Genes shown outside the circle are transcribed clockwise, and those inside are transcribed counterclockwise. The dark grey area in the inner circle indicates the CG content of the plastome.
Fig. 1.

Map of Cephalotaxus plastomes. Genes shown outside the circle are transcribed clockwise, and those inside are transcribed counterclockwise. The dark grey area in the inner circle indicates the CG content of the plastome.

Compared with Amentotaxus, Pseudotaxus, Taxus and Torreya, massive genic rearrangements were identified in Cephalotaxus by the progressive Mauve alignment of plastomes (Fig. 2). These variations led to Cephalotaxus plastomes being drastically different in gene arrangement from other Taxaceae genera. Nevertheless, Cephalotaxus plastomes exhibit high levels of sequence similarity at the species level. Plastome-wide mVISTA analyses (Fig. 3) detected a total of 2782 variations within the 139 245 alignment positions, representing 2 % variation in proportion among Cephalotaxus plastomes.

Multiple alignment resulted from Mauve showing genic arrangements detected in plastomes of Cephalotaxus and allied genera. Colour bars indicate syntenic blocks, and connecting lines indicate correspondence of blocks across plastomes.
Fig. 2.

Multiple alignment resulted from Mauve showing genic arrangements detected in plastomes of Cephalotaxus and allied genera. Colour bars indicate syntenic blocks, and connecting lines indicate correspondence of blocks across plastomes.

Alignment of Cephalotaxus plastomes using mVISTA, showing the percentages of sequence identity (y-axis).
Fig. 3.

Alignment of Cephalotaxus plastomes using mVISTA, showing the percentages of sequence identity (y-axis).

Phylogenetic reconstruction

The ML and BI analyses generated highly congruent tree topologies. The monophyly of Cephalotaxus (Cephalotaxaceae) was supported with 100 % bootstrap (BS) and 1.0 PP in ML and BI phylogenies. Its sister clade (Taxaceae) consisting of Taxus, Pseudotaxus, Torreya and Amentotaxus was fully supported (BS = 100 %, PP = 1.0) as well. Our plastome phylogenomic analyses recovered well-supported relationships among the Taxaceae genera: the four genera were grouped into two well-supported branches in the phylogenetic tree (BS = 100 %, PP = 1.0), in which the sister relationship between Taxus and Pseudotaxus, as well as between Torreya and Amentotaxus, received strong support (BS = 100 %, PP = 1.0). Within Cephalotaxus, C. oliveri was the earliest diverging species and sister to the clade consisting of the remaining species. Cephalotaxus harringtonii was sister to the clade including the other species, with this clade further containing two sister clades, one including four species (C. hainanensis, C. mannii, C. fortunei and C. griffithii), and the other including three species (C. sinensis, C. alpina and C. nana). Overall, our plastome-based phylogeny satisfactorily resolved the infrageneric relationships of Cephalotaxus: except for the C. fortunei and C. griffithii clade, all nodes within the genus are robust with a BS and PP of 100 % and 1.0, respectively (Fig. 4).

Phylogenetic relationships between Cephalotaxus and allied genera reconstructed by analyses of 81 plastid protein-coding genes using maximum likelihood (ML) and Bayesian inference (BI) methods. Numbers above branches indicated ML bootstrap percentage (BS) and BI posterior probability (PP).
Fig. 4.

Phylogenetic relationships between Cephalotaxus and allied genera reconstructed by analyses of 81 plastid protein-coding genes using maximum likelihood (ML) and Bayesian inference (BI) methods. Numbers above branches indicated ML bootstrap percentage (BS) and BI posterior probability (PP).

Divergence time estimation

The fossil-calibrated molecular dating (Fig. 5) indicated that the diversification of extant Cephalotaxus species initiated around 20.85 Mya [95 % highest posterior density (HPD): 39.30–8.52 Mya], which is the Oligocene/Miocene transition, with C. oliveri being the earliest diverging species in the genus. Subsequently, the speciation of C. harringtonii occurred at 8.02 Mya (95 % HPD: 15.22–3.22 Mya), in the late Miocene. The diversification of the remaining species was dated to 4.68 Mya (95 % HPD: 9.03–3.22 Mya), around the Miocene/Pliocene boundary, leading to the split of the clade including C. hainanensis, C. mannii, C. fortunei and C. griffithii, and the other consisting of C. sinensis, C. alpina and C. nana. The divergence of the aforementioned terminal species occurred in the Pleistocene. Additionally, the BEAST analyses assumed a median age of 144.47 Mya (95 % HPD: 160.16–126.11 Mya) for the crown age of Taxaceae. The crown ages of Amentotaxus, Torreya, and Taxus were dated to 2.06 Mya (95 % HPD: 4.39–0.56 Mya), 9.09 Mya (95 % HPD: 16.90–4.25 Mya) and 28.75 Mya (95 % HPD: 46.05–16.43 Mya), respectively. Similar to Cephalotaxus, the most extensive species divergence events in these genera, which are responsible for the occurrence of the majority of the extant species in each genus, occurred during late Miocene to Pleistocene (Fig. 5).

Divergence time estimation based on 81 plastid protein-coding genes. Numbers above and under the branches represent mean divergent ages and 95 % confidence interval of each node, respectively. Red arrows show the calibrating points for molecular dating. Divergence time and the timeline are indicated in million years ago (Mya).
Fig. 5.

Divergence time estimation based on 81 plastid protein-coding genes. Numbers above and under the branches represent mean divergent ages and 95 % confidence interval of each node, respectively. Red arrows show the calibrating points for molecular dating. Divergence time and the timeline are indicated in million years ago (Mya).

DISCUSSION

Phylogenetic inferences and taxonomic implications

Previous studies based on single- or multilocus DNA sequence data failed to recover an unambiguous relationship between Cephalotaxus and closely related genera (Chaw et al., 1997; Cheng et al., 2000; Quinn et al., 2002; Hao et al., 2008; Lu et al., 2014). Based on a large dataset of 81 protein-coding genes that contain more variable sites and parsimoniously informative variations than have previously been available (Supplementary data Table S2), and by far the most comprehensive taxonomic sampling (all extant Cephalotaxus species were included, and Taxus and Torreya were well represented), our plastome phylogenomic analyses not only reconstructed robust relationships of Cephalotaxus to allied genera, with high support for each intergeneric node (Fig. 4), but also provided a well-supported resolution of interspecific relationships within Cephalotaxus. Our results further confirmed that phylogenetic reconstruction based on plastome sequence data can effectively resolve historical problems in phylogenetically perplexing plant groups.

Although the sister relationship between Torreya and Amentotaxus has been commonly revealed by previous studies (Chaw et al., 1997; Cheng et al., 2000; Wang and Shu, 2000; Quinn et al., 2002; Hao et al., 2008; Lu et al., 2014; Majeed et al., 2018; Ran et al., 2018; X. Zhang et al., 2019), the position of these two genera remains unresolved. For instance, the phylogenetic analyses based on 18S ribosomal DNA (Chaw et al., 1997), ribosomal ITS regions (Cheng et al., 2000; Hao et al., 2008), LFY and NLY genes (Lu et al., 2014) and RNA sequencing (Majeed et al., 2018) revealed that they are closer to Cephalotaxus (Cephalotaxaceae) than to Taxaceae. However, phylogenetic reconstruction using plastid matK (Cheng et al., 2000), a combination of plastid matK, rbcL, trnL–F and psbA–trnH (Hao et al., 2008), transcriptomic data (Ran et al., 2018) and plastomes (X. Zhang et al., 2019; this study) yielded opposite results. Notably, those studies that reconstructed phylogeny based on nuclear sequence data, i.e. Chaw et al. (1997), Cheng et al. (2000), Lu et al. (2014) and Majeed et al. (2018), had limited taxonomic sampling from Cephalotaxaceae and Taxaceae. Accordingly, phylogenetic errors or uncertainty in the tree topology resulting from limited taxon sampling seem inevitable (Rokas and Carroll, 2005; Philippe et al., 2011). Although the ribosomal ITS sequences of a wider spectrum of taxon representing Cephalotaxaceae and Taxaceae were analysed by Hao et al. (2008), neither ML nor BI phylogeny was able to provide strong support to the sister relationship between Cephalotaxus and the Torreya + Amentotaxus clade.

With a much larger taxon sampling than has previously been available, plastome phylogenomics were capable of providing valuable and credible insights for elucidating the long-standing controversies in the relationships among Amentotaxus, Cephalotaxus, Pseudotaxus, Taxus and Torreya, and thus contribute to a clear-cut delineation between Cephalotaxaceae and Taxaceae. Briefly, Cephalotaxus (Cephalotaxaceae) was resolved as sister to Taxaceae; within Taxaceae, the sister relationships between Taxus and Pseudotaxus, as well as between Torreya and Amentotaxus, were strongly supported (Fig. 5). Although the plastomes of Cephalotaxus and allied genera are paternally inherited (Mogensen, 1996) and analysis of plastome DNA sequences can only represent the paternal histories of these genera, the relationships among these genera revealed by our data are highly congruent with that of transcriptome-based phylogeny (Ran et al., 2018). This implies that the previous disagreements regarding the systematic position of Amentotaxus and Torreya most probaby result from limited taxonomic sampling (e.g. Chaw et al., 1997; Lu et al., 2014; Majeed et al., 2018) or inadequate sequence variations (Hao et al., 2008) rather than conflicts between plastid and nuclear datasets. In view of this, the plastome-based phylogeny reconstructed in this study can credibly reflect the evolutionary relationships among these genera.

The relationships revealed by our data are consistent with some morphological characteristics: the two-ovulate bracts in the seed cones distinguish Cephalotaxus from the single-ovulate bracts of the rest of the genera (Xiao et al., 2008); among the Taxaceae genera, the sister relationship between Taxus and Pseudotaxus is supported by the seeds being imperfectly enveloped by a cup-like aril, in contrast to seeds completely enclosed by aril in the sister genera Torreya and Amentotaxus (Price, 1990). Remarkably, large-scale plastome structural rearrangements were observed in Cephalotaxaceae, making it distinct from Taxaceae in the organization of plastid genes (Fig. 2). In addition to morphological differences (Xiao et al., 2008), embryological investigations also identified clear distinctions between Cephalotaxaceae and Taxaceae (Singh, 1961). Collectively, these differences convincingly justify the distinctiveness of Cephalotaxaceae and Taxaceae, and thus they should be recognized as separate families. Therefore, our results offer plastome-based phylogenomic evidence to recognize Cephalotaxaceae as a monotypic family sister to Taxaceae (Cheng et al., 2000), and support the classification proposed by Janchen (1949), who divided Taxaceae into two tribes, Torreyeae (including Amentotaxus and Torreya) and Taxeae (including Pseudotaxus and Taxus).

The robust phylogeny reconstructed in this study also provides new insights into the species delineation in Cephalotaxus (Fig. 4). For instance, the well-supported sister relationship between C. wilsoniana and C. harringtonii was recovered, providing molecular evidence to support treating C. wilsoniana as either a conspecific variety (Fu, 1999; Farjon, 2010) or a synonym of the latter species (Lang et al., 2013; J. W. Zhang et al., 2019). In addition, the tree topologies indicated that C. alpina is phylogenetically distinct from C. fortunei, supporting that it should be recognized as a separate species (Fu, 1984; Lang et al., 2013; J. W. Zhang et al., 2019) rather than a conspecific variety under C. fortunei (Fu, 1999; Farjon, 2010). Nevertheless, as shown by our results, the treatment of C. mannii as a synonym of either C. harringtonii (Lang et al., 2013) or C. hainanensis (J. W. Zhang et al., 2019), as well as reducing C. sinensis to C. harringtonii (Lang et al., 2013; J. W. Zhang et al., 2019) are likely to be arbitrary.

Notably, none of the taxonomic revisions of Cephalotaxus is fully supported by our data, suggesting that the alpha taxonomy of the genus is still not well understood. Recently, DNA barcodes have been widely used for delineating species, particularly in taxonomically difficult plant groups. Because of their low effectiveness in discriminating recently diverged taxa, the standard DNA barcodes (i.e. rbcL, psbA–trnH and ITS) yield a weak performance in identifying Cephalotaxus species (Gao et al., 2015). Our results show that analysis of the plastome DNA sequence can significantly improve phylogenetic resolution in this genus. In addition, the great potential of plastome sequencing to accurately and reliably delineate species in taxonomically challenging plant genera, such as Araucaria (Ruhsam et al., 2015), Dendrobium (Zhu et al., 2019), Panax (Ji et al., 2019b), Paris (Ji et al., 2020) and Taxus (Fu et al., 2019), has been demonstrated in recent studies. Therefore, analyses of plastome DNA sequences by sampling multiple individuals from multiple congeneric species will help future taxonomic revision of Cephalotaxus through credible species delineation.

Recent species divergence in Cephalotaxus and allied genera

The divergence of Cephalotaxus species can be attributed to climate changes in the Cenozoic, in particular the establishment of monsoonal climates in East Asia and the Pleistocene climatic fluctuations. The initiation of the Asian monsoons in the Oligocene built a connection between forests from low to high latitudes of East Asia (Sun and Wang, 2005) and a humid climate in sub-tropical areas (Wan et al., 2007; Jacques et al., 2011). From that period onward, the forest corridor would allow the ancestral lineage of extant Cephalotaxus species to retreat to sub-tropical areas in response to the Neogene global cooling (Zachos et al., 2001, 2008). Interestingly, the earliest occurrence of Cephalotaxus in sub-tropical areas is a fossil species (C. ningmingensis) found in the Oligocene Ningming Formation of Guangxi, South China, which is most similar to extant C. oliveri in morphology (Shi et al., 2010). In addition, the early divergence of Cephalotaxus, resulting in the speciation of C. oliveri, was dated to the Oligocene/Miocene boundary (Fig. 5). Collectively, it is reasonable to hypothesize that the colonization of Cephalotaxus into low latitudes of East Asia, as well as the divergence of C. oliveri, could be attributed to the combination of the Neogene global climate cooling and the onset of the Asian monsoon.

The secondary divergence event in Cephalotaxus, involving the speciation of C. harringtonii, was dated at 8.02 Mya, around the late Miocene (Fig. 5). Since then, the intensification of Asian summer monsoons established a humid climate and caused a significant expansion of forests in East Asia (Sun and Wang, 2005; Yao et al., 2011). Such climatic and environmental shifts would drive the divergence of C. harringtonii and its northward migration to Japan and the Korean Peninsula. Moreover, the Middle Miocene fossil record of Cephalotaxus from Yunnan, south-west China (J. W. Zhang et al., 2019), suggests that the enhancement of monsoonal climates in East Asia played an important role in driving the westward dispersal of Cephalotaxus as well.

It is noteworthy that the most extensive species divergence events in Cephalotaxus, resulting in the divergence of the majority of the extant species, occurred in the Pleistocene (Fig. 5). During the Pleistocene, there were at least four major glaciations in East Asia (Shi et al., 2005). The Pleistocene glaciation/interglaciation cycles, which caused dramatic contraction/expansion of species ranges in the Northern Hemisphere (Axelrod et al., 1998; Qian and Ricklefs, 2000; Petit, 2003; Qiu et al., 2011), plus the increased complexity of topography in East Asia which might have blocked the regional gene flow and boosted vicariance (Wen et al., 2014; Favre et al., 2015), are believed to have triggered rapid speciation in many plant lineages in East Asia (Li and Li, 1997; Axelrod et al., 1998; Wu et al., 2005; Lu et al., 2018). Similarly, these events would have triggered significant species radiation in Cephalotaxus.

Both fossil evidence (Florin, 1958, 1963; He and Tao, 1997; Manchester et al., 2009; Leslie, 2018) and molecular dating consistently indicate that Cephalotaxus and all Taxaceae genera are ancient plant lineages. The antiquity of these genera can also be justified by the structural divergences among their plastomes. Despite high levels of similarity in gene content, massive genic rearrangements observed at the genus level (Fig. 2) suggest that plastomes are highly divergent in gene organization. The major differences are most likely to be associated with long-term divergent evolution of plastomes among these genera and thus accumulation of mutations and structural variations as inversions and relocation of genes. Nevertheless, divergence time estimation indicates that all extant species within the genus Cephalotaxus derived from recent divergence events that initiated around the Oligocene/Miocene transition and intensified in the Pleistocene. The young ages of extant Cephalotaxus species are consistent with the low levels of DNA sequence variations among their plastomes (Fig. 3). Recent diversification of extant Cephalotaxus species implies that they are not evolutionary relicts.

We found a similar scenario of species diversification in Amentotaxus, Taxus and Torreya: most extant species of these genera also originated from recent speciation events occurring in the late Miocene, Pliocene and Pleistocene, despite the origins of these genera occurring no later than the Paleogene according to the fossil record (Florin, 1958, 1963; Manchester et al., 2009; Leslie, 2018). Recent diversification of extant species has been observed in numerous ancient gymnosperm lineages. For instance, fossil-calibrated molecular phylogenies indicate that extant cycad species are not much older than approx. 12 million years although the cycad lineage is ancient (Nagalingum et al., 2011). Recent radiations have also been reported in many coniferous lineages in East Asia, such as Cupressus (Xu et al., 2010), Juniperus (Li et al., 2012), Larix (Wei and Wang, 2004), Picea (Ran et al., 2006, 2015) and Pinus (Liu et al., 2014; Hao et al., 2015; Liu et al., 2019). The scenario of ancient lineages possessing extremely high diversity of recently diverged species suggests that these gymnosperm genera may have suffered considerable extinction during their evolutionary histories (Nagalingum et al., 2011). Therefore, occurrence of such ancient lineages in a certain area does not equate to antiquity of the extant flora, as revealed by our results and previous studies.

Conclusions

This study is by far the most comprehensive taxonomic sampling, plastome-based phylogenomic inference of Cephalotaxus. The robust phylogeny reconstructed in this study provides a perspective on the phylogenetic placement of Cephalotaxus, and contributes to a clear-cut delineation between Cephalotaxaceae and Taxaceae. The well-supported resolution of interspecific relationships within Cephalotaxus provides new evidence to clarify some of the long-standing debates in species delineation. Our data reveal that the divergence of Cephalotaxus can be attributed to the establishment of monsoonal climates in East Asia and the Pleistocene climatic fluctuations, and the species richness in this genus is associated with recent divergence events occurring in the Pleistocene. Our findings reject the idea that extant Cephalotaxus species are evolutionary relicts (Fu, 1984; Wu et al., 2005). The evolutionary profiles of Cephalotaxus provide insightful knowledge for understanding the origin and evolution of the floristic paleo-endemism in East Asia.

SUPPLEMENTARY DATA

Supplementary data are available online at https://dbpia.nl.go.kr/aob and consist of the following. Table S1: plastome sequences downloaded from GenBank. Table S2: sequence characteristics of 81 protein-coding genes involved in the phylogenetic analyses. Table S3: summary of Illumina sequencing of Cephalotaxus species. Table S4: comparison of plastome features among Cephalotaxus species. Table S5: gene content of Cephalotaxus plastomes.

ACKNOWLEDGEMENTS

We thank Drs Xun Gong and Tao Su for providing some materials for this study, and to Wenyun Chen, Qiuping Tan, Sirong Yi and Qiliang Gan for their help with field work. The authors declare no conflict of interest related to this work.

FUNDING

This work was supported by the Major Program of National Natural Science Foundation of China (grant no. 31590823), the NSFC-Joint Foundation of Yunnan Province (grant no. U1802287) and a grant from Southeast Asia Biodiversity Research Institute, Chinese Academy of Science (grant no. Y4ZK111B01).

LITERATURE CITED

Axelrod
DI
.
1959
.
Poleward migration of early angiosperm flora
.
Science
130
:
203
207
.

Axelrod
DI
,
Al-Shehbaz
I
,
Raven
PH
.
1998
.
History of the modern flora of China
. In:
Zhang
AL
,
Wu
SG
, eds.
Floristic characteristics and diversity of East Asian plants
.
Beijing
:
China Higher Education Press
,
43
55
.

Barrett
CF
,
Specht
CD
,
Leebens-Mack
J
,
Stevenson
DW
,
Zomlefer
WB
,
Davis
JI
.
2014
.
Resolving ancient radiations: can complete plastid gene sets elucidate deep relationships among the tropical gingers (Zingiberales)?
Annals of Botany
113
:
119
133
.

Bouckaert
R
,
Heled
J
,
Kühnert
D
, et al.
2014
.
Beast 2: a software platform for Bayesian evolutionary analysis
.
PLoS Computational Biology
10
:
e1003537
.

Carlsen
MM
,
Fér
T
,
Roswitha
S
, et al.
2018
.
Resolving the rapid plant radiation of early diverging lineages in the tropical Zingiberales: pushing the limits of genomic data
.
Molecular Phylogenetics and Evolution
128
:
55
68
.

Chaw
SM
,
Zharkikh
A
,
Sung
HM
,
Lau
TC
,
Li
WH
.
1997
.
Molecular phylogeny of extant gymnosperms and seed plant evolution: analysis of nuclear 18S rRNA sequences
.
Molecular Biology and Evolution
14
:
56
68
.

Cheng
Y
,
Nicolson
RG
,
Tripp
K
,
Chaw
SM
.
2000
.
Phylogeny of taxaceae and cephalotaxaceae genera inferred from chloroplast matK gene and nuclear rDNA ITS region
.
Molecular Phylogenetics and Evolution
14
:
353
365
.

Christenhusz
M
,
Reveal
J
,
Farjon
A
,
Gardner
MF
,
Mill
RR
,
Chase
MW
.
2011
.
A new classification and linear sequence of extant gymnosperms
.
Phytotaxa
19
:
55
70
.

Darling
AE
,
Mau
B
,
Perna
NT
.
2010
.
ProgressiveMauve: multiple genome alignment with gene gain, loss and rearrangement
.
PLoS One
5
:
e11147
.

Dierckxsens
N
,
Mardulyn
P
,
Smits
G
.
2017
.
NOVOPlasty: de novo assembly of organelle genomes from whole genome data
.
Nucleic Acids Research
45
:
e18
.

Dong
W
,
Xu
C
,
Wu
P
, et al.
2018
.
Resolving the systematic positions of enigmatic taxa: manipulating the chloroplast genome data of Saxifragales
.
Molecular Phylogenetics and Evolution
126
:
321
330
.

Doyle
JJ
,
Doyle
JL
.
1987
.
A rapid DNA isolation procedure for small quantities of fresh leaf tissue
.
Phytochemistry Bulletin
19
:
11
15
.

Eichler
AW
.
1889
.
Coniferae
. In:
Engler
A
,
Prantl
K
, eds.
Die naturlichen pflanzenfamilien, Vol. 1
.
Leipzig, Germany
:
Engelmann
,
28
116
.

Endlicher
S
.
1847
.
Taxineae. Synopsis coniferarum
.
St. Gallen, Switzerland
:
Scheitlin & Zollikofer
,
228
244
.

Farjon
A
.
2010
.
A handbook of the world’s conifers.
Leiden & Boston
:
Brill
.

Favre
A
,
Päckert
M
,
Pauls
SU
, et al.
2015
.
The role of the uplift of the Qinghai–Tibetan Plateau for the evolution of Tibetan biotas
.
Biological Reviews
90
:
236
253
.

Florin
R
.
1948
.
On the morphology and relationships on the Taxaceae
.
Botanical Gazette
110
:
31
39
.

Florin
R
.
1954
.
The female reproductive organs of conifers and taxads
.
Biological Reviews
29
:
367
389
.

Florin
R
.
1958
.
On jurassic taxads and conifers from north-Western Europe and Eastern Greenland
.
Uppsala
:
Almqvist & Wiksells Boktryckeri
.

Florin
R
.
1963
.
The distribution of conifer and taxad genera in time and space
.
Acta Horti Bergiani
20
:
121
312
.

Fu
CN
,
Wu
CS
,
Ye
LJ
, et al.
2019
.
Prevalence of isomeric plastomes and effectiveness of plastome super-barcodes in yews (Taxus) worldwide
.
Scientific Reports
9
:
2773
.

Fu
LG
.
1984
.
A study on the genus Cephalotaxus Sieb. et Zucc
.
Acta Phytotaxonomica Sinica
4
:
277
288
.

Fu
LG
,
Li
N
,
Mill
RR
.
1999
.
Cephalotaxus Sieb. et Zucc
. In:
Wu
ZY
,
Raven
PH
, eds.
Flora of China 4
.
St. Louis, MO
:
Missouri Botanical Garden Press
and
Beijing
:
Science Press,
85
88
.

Gao
LM
,
Li
DZ
,
Liu
J
.
2015
.
DNA barcoding for identification of Cephalotaxus and the discovery of new species
.
Genome
58
:
219
219
.

Ghimire
B
,
Heo
K
.
2014
.
Cladistic analysis of Taxaceae s. l
.
Plant Systematics
300
:
217
223
.

Givulescu
R
.
1973
.
Die fossilen Koniferen des Fundortes Chiuzbaia
.
Memorii Institutul de Geologie si Geofizica
19
:
31
34
.

Hao
DC
,
Xiao
PG
,
Huang
BL
,
Ge
GB
,
Yang
L
.
2008
.
Interspecific relationships and origins of Taxaceae and Cephalotaxaceae revealed by partitioned Bayesian analyses of chloroplast and nuclear DNA sequences
.
Plant Systematics and Evolution
276
:
89
104
.

Hao
ZZ
,
Liu
YY
,
Nazaire
M
,
Wei
XX
,
Wang
XQ
.
2015
.
Molecular phylogenetics and evolutionary history of sect. Quinquefoliae (Pinus): implications for Northern Hemisphere biogeography
.
Molecular Phylogenetics
87
:
65
79
.

Hart
JA
.
1987
.
A cladistic analysis of conifers: preliminary results
.
Journal of the Arnold Arboretum
26
:
9
307
.

He
CX
,
Tao
JR
.
1997
.
A study on the Eocene flora in Yilan County, Heilongjiang
.
Acta Phytotaxonomica Sinica
35
:
249
256
.

Heckenhauer
J
,
Paun
O
,
Chase
MW
,
Ashton
PS
,
Kamariah
AS
,
Samuel
R
.
2019
.
Molecular phylogenomics of the tribe Shoreeae (Dipterocarpaceae) using whole plastid genomes
.
Annals of Botany
20
:
1
9
.

Jacques
FM
,
Guo
SX
,
Su
T
, et al.
2011
.
Quantitative reconstruction of the Late Miocene monsoon climates of southwest China: a case study of the Lincang flora from Yunnan Province
.
Palaeogeography, Palaeoclimatology, Palaeoecology
304
:
318
327
.

Janchen
E
.
1949
.
Das System der Konifern
.
Akademie der Wissenschaften in Wien. Sitzungsberichte
158
:
155
262
.

Ji
Y
,
Yang
L
,
Chase
MW
, et al.
2019
a.
Plastome phylogenomics, biogeography, and clade diversification of Paris (Melanthiaceae)
.
BMC Plant Biology
19
:
543
.

Ji
Y
,
Liu
C
,
Yang
Z
, et al.
2019
b.
Testing and using complete plastomes and ribosomal DNA sequences as the next generation DNA barcodes in Panax (Araliaceae)
.
Molecular Ecology Resources
19
:
1333
1345
.

Ji
Y
,
Liu
C
,
Yang
J
,
Jin
L
,
Yang
Z
,
Yang
J-B
.
2020
.
Ultra-Barcoding discovers a cryptic species in Paris yunnanensis (Melanthiaceae), a medicinally important plant
.
Frontiers in Plant Science
11
:
411
.

Katoh
K
,
Standley
DM
.
2013
.
MAFFT multiple sequence alignment software version 7: improvements in performance and usability
.
Molecular Biology and Evolution
30
:
772
780
.

Kvaček
Z
,
Rember
WC
.
2000
.
Shared Miocene conifers of the Clarkia flora and Europe
.
Acta Universitatis Carolinae Geologica
44
:
75
85
.

Kvaček
Z
,
Walther
Η
.
1998
.
The Oligocene volcanic flora of Kundratice near Litoměřice České středohoří Vulcanic Complex (Czech Republic) – a review
.
Acta Musei Nationalis Pragae, Series B, Historia Naturalis Series В
54
:
1
42
.

Lanfear
R
,
Frandsen
PB
,
Wright
AM
,
Senfeld
T
,
Calcott
B
.
2017
.
PartitionFinder 2: new methods for selecting partitioned models of evolution for molecular and morphological phylogenetic analyses
.
Molecular Biology
34
:
772
773
.

Lang
XD
,
Su
JR
,
Lu
SG
,
Zhang
ZJ
.
2013
.
A taxonomic revision of the genus Cephalotaxus (Taxaceae)
.
Phytotaxa
84
:
1
24
.

Leslie
AB
,
Beaulieu
J
,
Holman
G
, et al.
2018
.
An overview of extant conifer evolution from the perspective of the fossil record
.
American Journal of Botany
105
:
1531
1544
.

Li
HT
,
Yi
TS
,
Gao
LM
, et al.
2019
.
Origin of angiosperms and the puzzle of the Jurassic gap
.
Nature Plants
5
:
461
470
.

Li
XW
,
Li
J
.
1997
.
The Tanaka–Kaiyong line – an important floristic line for the study of the flora of East Asia
.
Annals of the Missouri Botanical Garden
84
:
888
892
.

Li
ZH
,
Zou
JB
,
Mao
KS
, et al.
2012
.
Population genetic evidence for complex evolutionary histories of four high altitude juniper species in the Qinghai–Tibetan Plateau
.
Evolution
66
:
831
845
.

Liu
L
,
Hao
ZZ
,
Liu
YY
,
Wei
XX
,
Cun
YZ
,
Wang
XQ
.
2014
.
Phylogeography of Pinus armandii and its relatives: heterogeneous contributions of geography and climate changes to the genetic differentiation and diversification of Chinese white pines
.
PLoS One
9
:
e85920
.

Liu
YY
,
Jin
WT
,
Wei
XX
,
Wang
XQ
.
2019
.
Cryptic speciation in the Chinese white pine (Pinus armandii): implications for the high species diversity of conifers in the Hengduan Mountains, a global biodiversity hotspot
.
Molecular Phylogenetics and Evolution
138
:
114
125
.

Lu
LM
,
Mao
LF
,
Yang
T
, et al.
2018
.
Evolutionary history of the angiosperm flora of China
.
Nature
554
:
234
238
.

Lu
Y
,
Ran
JH
,
Guo
DM
,
Yang
ZY
,
Wang
XQ
.
2014
.
Phylogeny and divergence times of gymnosperms inferred from single-copy nuclear genes
.
Plos One
9
:
e107679
.

Mai
DH
,
Walther
H
.
1978
.
Die Floren der Haselbacher Serie im Weißelster-Becken (Bezirk Leipzig, DDR)
.
Abhandlungen des Staatlichen Museums für Mineralogie und Geologie zu Dresden
28
:
1
200
.

Majeed
A
,
Singh
A
,
Choudhary
S
,
Bhardwaj
P
.
2018
.
RNAseq-based phylogenetic reconstruction of Taxaceae and Cephalotaxaceae
.
Cladistics
35
:
461
468
.

Manchester
SR
,
Chen
ZD
,
Lu
AM
,
Uemura
K
.
2009
.
Eastern Asian endemic seed plant genera and their paleogeographic history throughout the Northern Hemisphere
.
Journal of Systematics and Evolution
47
:
1
42
.

Mayor
C
,
Brudno
M
,
Schwartz
JR
, et al.
2000
.
VISTA: visualizing global DNA sequence alignments of arbitrary length
.
Bioinformatics
16
:
1046
1047
.

Meller
B
.
1998
.
Systematish-taxonomishe Untersuchungen von Karpo-Taphocoenosen des Köflach-Voitsberger Braunkohlenrevieres (Steiermark, Östereich; Untermiozän) und ihre paläoökologische Bedeutung
.
Jahrbuch der Geologischen Bundesanstalt Wien
140
:
497
655
.

Miki
S
.
1958
.
Gymnosperms in Japan, with special reference to its remains
.
Journal of the Institute of Polytechnics, Osaka City University, Series D
9
:
221
272
.

Mogensen
HL
.
1996
.
The hows and whys of cytoplasmic inheritance in seed plants
.
American Journal of Botany
83
:
221
272
.

Nagalingum
NS
,
Marshall
CR
,
Quental
TB
,
Rai
HS
,
Little
DP
,
Mathews
S
.
2011
.
Recent synchronous radiation of a living fossil
.
Science
334
:
796
799
.

Neger
FW
.
1907
.
Cephalotaxaceae
. In:
Göschen
GJ
, ed.
Die Nadelhölzer (Koniferen) und Übrigen
Gymnospermen.
Leipzig, Germany
:
De Gruyter
,
23
.

Parks
M
,
Cronn
R
,
Liston
A
.
2009
.
Increasing phylogenetic resolution at low taxonomic levels using massively parallel sequencing of chloroplast genomes
.
BMC Biology
7
:
84
.

Patel
RK
,
Jain
M
.
2012
.
NGS QC Toolkit: a toolkit for quality control of next generation sequencing data
.
PLoS One
7
:
e30619
.

Petit
RJ
,
Aguinagalde
I
,
de Beaulieu
JL
, et al.
2003
.
Glacial refugia: hotspots but not melting pots of genetic diversity
.
Science
300
:
1563
1565
.

Philippe
H
,
Brinkmann
H
,
Lavrov
DV
, et al.
2011
.
Resolving difficult phylogenetic questions: why more sequences are not enough
.
PLoS Biology
9
:
e1000602
.

Pilger
R
.
1903
.
Taxaceae
. In:
Engler
A
, ed.
Das Pflanzenreich 5 (Heft 18)
, 4th edn.
Leipzig, Germany
:
Wilhelm Engelmann
,
99
105
.

Posada
D
,
Buckley
TR
.
2004
.
Model selection and model averaging in phylogenetics: advantages of Akaike information criterion and Bayesian approaches over likelihood ratio tests
.
Systematic Biology
53
:
793
808
.

Posada
D
,
Crandall
KA
.
1998
.
MODELTEST: testing the model of DNA substitution
.
Bioinformatics
14
:
817
818
.

Price
RA
.
1990
.
The genera of Taxaceae in the southeastern United States
.
Journal of the Arnold Arboretum
71
:
69
91
.

Qian
H
,
Ricklefs
RE
.
2000
.
Large-scale processes and the Asian bias in species diversity of temperate plants
.
Nature
407
:
180
182
.

Qiu
YX
,
Fu
CX
,
Comes
HP
.
2011
.
Plant molecular phylogeography in China and adjacent regions: tracing the genetic imprints of Quaternary climate and environmental change in the world’s most diverse temperate flora
.
Molecular Phylogenetics and Evolution
59
:
225
244
.

Quinn
CJ
,
Price
RA
,
Gadek
PA
.
2002
.
Familial concepts and relationships in the conifer based on rbcL and matK sequence comparisons
.
Kew Bulletin
57
:
513
531
.

Rambaut
A
,
Drummond
AJ
,
Xie
D
,
Baele
G
,
Suchard
MA
.
2018
.
Posterior summarisation in Bayesian phylogenetics using Tracer 1.7
.
Systematic Biology
5
:
901
904
.

Ran
JH
,
Wei
XX
,
Wang
XQ
.
2006
.
Molecular phylogeny and biogeography of Picea (Pinaceae): implications for phylogeographical studies using cytoplasmic haplotypes
.
Molecular Phylogenetics and Evolution
41
:
405
419
.

Ran
JH
,
Shen
TT
,
Liu
WJ
,
Wang
PP
,
Wang
XQ
.
2015
.
Mitochondrial introgression and complex biogeographic history of the genus Picea
.
Molecular Phylogenetics and Evolution
93
:
63
76
.

Ran
JH
,
Shen
TT
,
Wang
MM
,
Wang
XQ
.
2018
.
Phylogenomics resolves the deep phylogeny of seed plants and indicates partial convergent or homoplastic evolution between Gnetales and angiosperms
.
Proceedings of the Royal Society B: Biological Sciences
285
:
20181012
.

Rokas
A
,
Carroll
SB
.
2005
.
More genes or more taxa? The relative contribution of gene number and taxon number to phylogenetic accuracy
.
Molecular Biology and Evolution
22
:
1337
1344
.

Ronquist
F
,
Huelsenbeck
JP
.
2003
.
MrBayes 3: Bayesian phylogenetic inference under mixed models
.
Bioinformatics
19
:
1572
1574
.

Ruhsam
M
,
Rai
HS
,
Mathews
S
, et al.
2015
.
Does complete plastid genome sequencing improve species discrimination and phylogenetic resolution in Araucaria?
Molecular Ecology Resources
15
:
1067
1078
.

Schattner
P
,
Brooks
AN
,
Lowe
TM
.
2005
.
The tRNAscan-SE, snoscan and snoGPS web servers for the detection of tRNAs and snoRNAs
.
Nucleic Acids Research
33
:
W686
W689
.

Shi
G
,
Zhou
Z
,
Xie
Z
.
2010
.
A new Cephalotaxus and associated epiphyllous fungi from the Oligocene of Guangxi, South China
.
Review of Palaeobotany and Palynology
161
:
179
195
.

Shi
YF
,
Cui
ZJ
,
Su
Z
.
2005
.
The Quaternary glaciations and environmental variations in China
.
Shijiangzhuang
:
Hebei Science and Technology Publishing House
.

Simpson
MG
,
Guilliams
CM
,
Hasenstab-Lehman
KE
,
Mabry
ME
,
Ripma
L
.
2017
.
Phylogeny of the popcorn flowers: use of genome skimming to evaluate monophyly and interrelationships in subtribe Amsinckiinae (Boraginaceae)
.
Taxon
66
:
1406
1420
.

Singh
H
.
1961
.
The life history and systematic position of Cephalotaxus drupaceae Sieb. et Zucc
.
Phytomorphology
11
:
153
197
.

Stamatakis
A
.
2006
.
RAxML-VI-HPC: maximum likelihood-based phylogenetic analysis with thousands of taxa and mixed models
.
Bioinformatics
22
:
2688
2690
.

Stefanoviac
S
,
Jager
M
,
Deutsch
J
,
Broutin
J
,
Masselot
M
.
1998
.
Phylogenetic relationships of conifers inferred from partial 28S rRNA gene sequences
.
American Journal of Botany
85
:
688
.

Stull
GW
,
Dunod
SR
,
Soltis
DE
,
Soltis
PS
.
2015
.
Resolving basal lamiid phylogeny and the circumscription of Icacinaceae with a plastome-scale data set
.
American Journal of Botany
102
:
1794
1813
.

Sun
XJ
,
Wang
PX
.
2005
.
How old is the Asian monsoon system? Palaeobotanical records from China
.
Palaeogeography, Palaeoclimatology, Palaeoecology
222
:
181
222
.

Takhtajan
AL
.
1953
.
Phylogenetic principles of the system of higher plants
.
The Botanical Review
19
:
1
45
.

Tripp
KE
.
1995
.
Cephalotaxus: the plum yews
.
Arnoldia
55
:
24
40
.

Uribe-Convers
S
,
Carlsen
MM
,
Lagomarsino
LP
,
Muchhala
N
.
2017
.
Phylogenetic relationships of Burmeistera (Campanulaceae: Lobelioideae): combining whole plastome with targeted loci data in a recent radiation
.
Molecular Phylogenetics and Evolution
107
:
551
563
.

Van Tieghem
MP
.
1891
.
Structure et affinities des Cephalotaxus
.
Bulletin de la Societe Botanique de France
38
:
184
190
.

Walther
H
,
Kvaček
Z
.
2007
.
The Oligocene flora of Seifhennersdorf
.
Acta Musei Nationalis Pragae, Series B, Historia Naturalis
63
:
85
174
.

Wan
SM
,
Li
AC
,
Clift
PD
,
Stut
JBW
.
2007
.
Development of the East Asian monsoon: mineralogical and sedimentologic records in the northern South China Sea since 20 Ma
.
Palaeogeography, Palaeoclimatology, Palaeoecology
254
:
561
582
.

Wang
X
,
Shu
Y
.
2000
.
Chloroplast matK gene phylogeny of Taxaceae and Cephalotaxaceae, with additional reference to the systematic position of Nageia
.
Acta Phytotaxonomica Sinica
38
:
201
210
.

Wei
XX
,
Wang
XQ
.
2004
.
Recolonization and radiation in Larix (Pinaceae): evidence from nuclear ribosomal DNA paralogues
.
Molecular Ecology
13
:
3115
3123
.

Wen
J
,
Zhang
JQ
,
Nie
ZL
,
Zhong
Y
,
Sun
H
.
2014
.
Evolutionary diversifications of plants on the Qinghai–Tibetan Plateau
.
Frontiers in Genetics
5
:
4
.

Whitfield
JB
,
Lockhart
PJ
.
2007
.
Deciphering ancient rapid radiations
.
Trends in Ecology and Evolution
22
:
258
265
.

Wilde
V
.
1989
.
Untersuchungen zur Systematik der Blattreste aus dem Mitteleozän der Grube Messel bei Darmstadt (Hessen, Bundesrepublik Deutschland)
.
Courier Forschungstinstut Senckenberg
115
:
1
213
.

Wu
ZY
,
Sun
H
,
Zhou
ZK
,
Peng
H
,
Li
DZ
.
2005
.
Origin and differentiation of endemism in the flora of China
.
Acta Botanica Yunnanica
27
:
577
604
.

Wyman
SK
,
Jansen
RK
,
Boore
JL
.
2004
.
Automatic annotation of organellar genomes with DOGMA
.
Bioinformatics
20
:
3252
3255
.

Xiao
PG
,
Huang
B
,
Ge
GB
,
Yang
L
.
2008
.
Interspecific relationships and origins of Taxaceae and Cephalotaxaceae revealed by partitioned Bayesian analyses of chloroplast and nuclear DNA sequences
.
Plant Systematics
276
:
89
104
.

Xu
T
,
Abbott
RJ
,
Milne
RI
, et al.
2010
.
Phylogeography and allopatric divergence of cypress species (Cupressus L.) in the Qinghai–Tibetan Plateau and adjacent regions
.
BMC Biology
10
:
194
.

Xu
XH
,
Sun
BN
,
Yan
DF
,
Wang
J
,
Dong
C
.
2015
.
A Taxus leafy branch with attached ovules from the Lower Cretaceous of Inner Mongolia, North China
.
Cretaceous Research
54
:
266
282
.

Yang
Y
,
Wang
ZH
,
Xu
XT
.
2017
.
Taxonomy and distribution of global gymnosperms.
Shanghai, China
:
Shanghai Science and Technology Press
,
1223
.

Yao
YF
,
Bruch
AA
,
Mosbrugger
V
,
Li
CS
.
2011
.
Quantitative reconstruction of Miocene climate patterns and evolution in Southern China based on plant fossils
.
Palaeogeography, Palaeoclimatology, Palaeoecology,
304
:
291
307
.

Zachos
JC
,
Pagani
M
,
Sloan
L
,
Thomas
E
,
Billups
K
.
2001
.
Trends, rhythms, and aberrations in global climate 65 Ma to present
.
Science
292
:
686
693
.

Zachos
JC
,
Dickens
GR
,
Zeebe
RE
.
2008
.
An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics
.
Nature
451
:
279
283
.

Zhang
JW
,
D’Rozario
A
,
Liang
XQ
,
Zhou
ZK
.
2019
.
Middle Miocene Cephalotaxus (Taxaceae) from Yunnan, Southwest China, and its implications to taxonomy and evolution of the genus
.
Palaeoworld
28
:
381
402
.

Zhang
SD
,
Jin
JJ
,
Chen
SY
, et al.
2017
.
Diversification of Rosaceae since the Late Cretaceous based on plastid phylogenomics
.
New Phytologist
214
:
1355
1367
.

Zhang
X
,
Zhang
HJ
,
Landis
JB
, et al.
2019
.
Plastome phylogenomic analysis of Torreya (Taxaceae)
.
Journal of Systematics and Evolution
57
:
607
615
.

Zhu
S
,
Niu
Z
,
Xue
Q
,
Wang
H
,
Xie
X
,
Ding
X
.
2019
.
Accurate authentication of Dendrobium officinale and its closely related species by comparative analysis of complete plastomes
.
Acta Pharmaceutica Sinica B
8
:
969
980
.

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse, distribution, and reproduction in any medium, provided the original work is properly cited.