Europe PMC
Do data resources managed by EMBL-EBI and our collaborators make a difference to your work?
If so, please take 10 minutes to fill in our survey, and help us make the case for why sustaining open data resources is critical for life sciences research.

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Background and aims

The largest genus of Salicaceae sensu lato, Salix, has been shown to consist of two main clades: clade Salix, in which species have XY sex-determination systems (SDSs) on chromosome 7, and clade Vetrix including species with ZW SDSs on chromosome 15. Here, we test the utility of whole genome re-sequencing (WGR) for phylogenomic reconstructions of willows to infer changes between different SDSs.

Methods

We used more than 1 TB of WGR data from 70 Salix taxa to ascertain single nucleotide polymorphisms on the autosomes, the sex-linked regions (SLRs) and the chloroplast genomes, for phylogenetic and species tree analyses. To avoid bias, we chose reference genomes from both groups, Salix dunnii from clade Salix and S. purpurea from clade Vetrix.

Key results

Two main largely congruent groups were recovered: the paraphyletic Salix grade and the Vetrix clade. The autosome dataset trees resolved four subclades (C1-C4) in Vetrix. C1 and C2 comprise species from the Hengduan Mountains and adjacent areas and from Eurasia, respectively. Section Longifoliae (C3) grouped within the Vetrix clade but fell into the Salix clade in trees based on the chloroplast dataset analysis. Salix triandra from Eurasia (C4) was revealed as sister to the remaining species of clade Vetrix. In Salix, the polyploid group C5 is paraphyletic to clade Vetrix and subclade C6 is consistent with Argus's subgenus Protitea. Chloroplast datasets separated both Vetrix and Salix as monophyletic, and yielded C5 embedded within Salix. Using only diploid species, both the SLR and autosomal datasets yielded trees with Vetrix and Salix as well-supported clades.

Conclusion

WGR data are useful for phylogenomic analyses of willows. The different SDSs may contribute to the isolation of the two major groups, but the reproductive barrier between them needs to be studied.

Free full text 


Logo of annbotAboutAuthor GuidelinesEditorial BoardAnnals of Botany
Ann Bot. 2022 Apr 1; 129(4): 485–498.
Published online 2022 Mar 23. https://doi.org/10.1093/aob/mcac012
PMCID: PMC8944726
PMID: 35134824

The phylogeny of Salix revealed by whole genome re-sequencing suggests different sex-determination systems in major groups of the genus

Associated Data

Supplementary Materials

Abstract

Background and Aims

The largest genus of Salicaceae sensu lato, Salix, has been shown to consist of two main clades: clade Salix, in which species have XY sex-determination systems (SDSs) on chromosome 7, and clade Vetrix including species with ZW SDSs on chromosome 15. Here, we test the utility of whole genome re-sequencing (WGR) for phylogenomic reconstructions of willows to infer changes between different SDSs.

Methods

We used more than 1 TB of WGR data from 70 Salix taxa to ascertain single nucleotide polymorphisms on the autosomes, the sex-linked regions (SLRs) and the chloroplast genomes, for phylogenetic and species tree analyses. To avoid bias, we chose reference genomes from both groups, Salix dunnii from clade Salix and S. purpurea from clade Vetrix.

Key Results

Two main largely congruent groups were recovered: the paraphyletic Salix grade and the Vetrix clade. The autosome dataset trees resolved four subclades (C1–C4) in Vetrix. C1 and C2 comprise species from the Hengduan Mountains and adjacent areas and from Eurasia, respectively. Section Longifoliae (C3) grouped within the Vetrix clade but fell into the Salix clade in trees based on the chloroplast dataset analysis. Salix triandra from Eurasia (C4) was revealed as sister to the remaining species of clade Vetrix. In Salix, the polyploid group C5 is paraphyletic to clade Vetrix and subclade C6 is consistent with Argus’s subgenus Protitea. Chloroplast datasets separated both Vetrix and Salix as monophyletic, and yielded C5 embedded within Salix. Using only diploid species, both the SLR and autosomal datasets yielded trees with Vetrix and Salix as well-supported clades.

Conclusion

WGR data are useful for phylogenomic analyses of willows. The different SDSs may contribute to the isolation of the two major groups, but the reproductive barrier between them needs to be studied.

Keywords: Salix, phylogeny, sex determination, whole genome re-sequencing, chloroplast dataset, willows

Introduction

Salicaceae sensu lato (s.l.) includes over 50 genera, with ~1000 species of woody trees and shrubs (Chase et al., 2002; Li et al., 2019). Salix L. (willows) is the largest genus (Skvortsov, 1999; Fang et al., 1999; Ohashi et al., 2006), and includes ~450 species mainly distributed in the Northern Hemisphere (reviewed in Argus, 1997; Skvortsov, 1999; He et al., 2021a). However, species with valuable biological features have been widely introduced and cultivated in various locations around the world (Isebrands and Richardson, 2014).

Reproduction by separate sexes (dioecy), reduced flowers, polyploidization and frequent natural hybridization, with often wide ranges of intraspecific phenotypic and genotypic variation, all create problems for willow classification (Argus, 1997; Skvortsov, 1999; Fang et al., 1999). Molecular evidence in the first decade of the 21st century proved Salix to be a monophyletic group, and significant progress in genus delimitation has been achieved (Supplementary Data Table S1).

Subgeneric classification of Salix has, however, remained in a process of endless revision (Supplementary Data Table S1). Skvortsov (1968, translated into English in Skvortsov, 1999), recognized three subgenera in Eurasia, Salix, Chamaetia and Vetrix. However, he made no taxonomic decision on species from other continents. He admitted that the separation between the subgenera Chamaetia and Vetrix was not clear. These subgenera are more closely related to each other than to Salix s.l. (sensu Skvortsov), which shows ‘primitive’ morphological features (Barkalov and Kozyrenko, 2014; Cronk et al., 2015; Wu et al., 2015) typically found in Populus. Dorn (1976), based on morphological characteristics of American willows, accepted only the subgenera Salix and Vetrix.Argus (1997) conducted a morphological cladistic analysis and classified the North American species into four subgenera: Salix, Longifoliae, Chamaetia and Vetrix. Later, Argus (2010) accepted five subgenera in the Flora of North America: Protitea, Longifoliae, Salix, Chamaetia and Vetrix.

Hardig et al. (2010) used matK chloroplast markers and ribosomal DNA internal transcribed spacer (ITS) sequences, generally supporting Argus’s subgenera. However, Chen et al. (2010) revealed the two subgenera Chamaetia and Vetrix to be monophyletic based on three plastid markers. Abdollahzadeh et al. (2011) considered that all of these subgenera were non-monophyletic except Longifoliae. Using external transcribed spacer (ETS) and ITS sequences of nuclear ribosomal DNA and four plastid markers, Wu et al. (2015) supported the merging of the subgenera Chamaetia and Vetrix, with sect. Amygdalinae as sister to it. Lauron-Moreau et al. (2015a) proposed to subdivide Salix into two subgenera (Salix and Vetrix), but in a corrected version (Lauron-Moreau et al., 2015b) four clades were recognized, basically consistent with subgenera Protitea, Salix, Longifoliae and Vetrix (including Chamaetia), respectively. Other molecular studies around the same time mostly recognized two major clades within Salix: one which is composed of species from subgenera Salix, Longifoliae and Protitea, and the other including species of subgenera Chamaetia and Vetrix, along with sect. Amygdalinae, and representatives of the formerly recognized segregated genera Chosenia (Salix arbutifolia) and Toisusu (Salix cardiophylla) (Wu et al., 2015; Zhang et al., 2018b). Restriction-site associated DNA (RAD) sequencing has been applied to estimate the phylogeny of willows, whereby in particular Wagner et al. (2018, 2020, 2021a) confirmed the monophyly of Chamaetia and Vetrix, suggesting they be treated as the Chamaetia/Vetrix clade, but excluding sect. Amygdalinae. He et al. (2021a) discussed adaptive evolution patterns of some Chamaetia/Vetrix species and their radiation in the Hengduan Mountains, showing subdivision within the clade into the Hengduan and Eurasian subclades. Recently, chloroplast genomes of 32 species confirmed the monophyly of three well-supported clades that are each separated on long branches: Chamaetia/Vetrix, subg. Salix, and in between the ‘Amygdalinae’ clade with Salix triandra (Wagner et al., 2021b).

The emergence and development of high-throughput sequencing technologies has provided new approaches (Morey et al., 2013). In particular, it has become possible to sequence large genomes at low cost with relative confidence in the data quality (Shendure et al., 2017), and whole genome sequencing (WGS) has become widely used (Unamba et al., 2015), and further extended to whole genome re-sequencing (WGR) (Shendure et al., 2017). This method can now be used for phylogenomic studies (e.g. Malmstrøm et al., 2017; Ma et al., 2018; Olofsson et al., 2019).

Compared to the nuclear genome, the chloroplast genome is considerably smaller, varying among plants in size and coding genes (Daniell et al., 2016). Moreover, the substitution rate of the chloroplast genome is low, which makes it potentially a useful resource for molecular phylogenetic studies (Raubeson and Jansen, 2005). The disadvantage is the low resolution of plastid phylogenies and the predominant maternal inheritance of the chloroplast genome. Phylogenetic incongruence with nuclear data may appear due to the maternal or biparental mode of transmission (Stull et al., 2020). Offspring resulting from interspecific hybridization inherit nuclear genes from both parents. However, as shown by Zhang and Liu (2003), since Salix has maternal inheritance, the chloroplast genome is expected to remain largely identical with that of the female parent. Several studies have applied chloroplast genomes to willow species to resolve phylogenetic issues. Zhang et al. (2018b) used whole chloroplast genome sequences of 42 members of Salicaceae s.l. mainly for divergence time estimation; nevertheless, two major clades within Salix, mainly comprising species of Chamaetia/Vetrix and Salix, respectively, were recognized. Li et al. (2019) used chloroplast genomes of Salix interior along with chloroplast genomes of 23 species from Salicaceae s.l. to reconstruct intrageneric relationships within the family. Wagner et al. (2021b) confirmed that chloroplast genomes can separate large clades, but that chloroplast genome evolution at the species level is shaped by low divergence, reticulate evolution and homoplasy. Thus, despite low resolution and low support values within clades, chloroplast genome data can be successfully used for investigation of the major clades of Salix.

Hybridization and sex determination

Salix is of interest for the study of sex-determination systems (SDSs), and these may relate to the phylogenetic relationships in the following way, making it important to clarify relationships among willows. Populus and Salix are sister genera in Salicaceae, and both are dioecious. It has therefore been suggested that this state was present in the common ancestor of these genera (Dai et al., 2014) before they diverged from each other about 40–45 Mya (Boucher et al., 2003; Wu et al., 2015). Despite the long period during which dioecy could have existed, the chromosomes carrying the sex-determination region (SDR) are homomorphic (van Buijtenen and Einspahr, 1959). In both genera the SDR systems are based on a single factor located on either the female (W) chromosome or the male (Y) chromosome (Renner and Müller, 2021). This factor is either female-specific expressed on the W chromosome or dominantly repressed by the male Y chromosome. Changes in the position of the SDR can occur relatively easily by translocation or new mutations (Renner and Müller, 2021). It is therefore possible that the SDR has evolved to some extent independently in these two genera, and that turnover events could have occurred, in which an established sex-determining system was replaced by a different one. Indeed, both male and female heterogamety are now known within the genus Salix. Studies of SDSs have mainly been based on members of subg. Vetrix, including Salix polyclona, S. suchowensis, S. viminalis, S. purpurea and S. triandra, which have female heterogamety (ZW) with physically extensive sex-linked regions (SLRs) on chromosome 15 in all species (reviewed in He et al., 2021b; Table 1). However, recent studies have demonstrated male heterogamety (XY) and SLRs on chromosome 7 in two species of subg. Salix, namely S. dunnii (He et al., 2021b) and S. nigra (Sanderson et al., 2021). Thus, the different SDSs may support a biologically important subdivision of Salix that could be used as a character in subgeneric classification (although sex determination has so far been studied in only a few species of Salix).

Table 1.

Summary of current information about sex determination systems in Salix, adapted from He et al. (2021b)

SpeciesMale or female heterogametyChromosome carrying the sex-determining locusReferences
Salix clade
 S. dunnii Male (XX/XY)7 He et al. (2021b)
 S. nigra Male (XX/XY)7 Sanderson et al. (2021)
Vetrix clade
 S. triandra Female (ZW/ZZ)15 Li et al. (2020)
 S. purpurea Female (ZW/ZZ)15 Zhou et al. (2020)
 S. suchowensis Female (ZW/ZZ)15 Hou et al. (2015)
 S. viminalis Female (ZW/ZZ)15 Almeida et al. (2020)
 S. polyclona Female (ZW/ZZ)15L. He et al. (in prep.)

Recent studies have tended to decrease the number of subgenera, predominantly agreeing on the recognition of only two major clades, the Salix clade consisting of subg. Salix, Longifoliae and Protitea, and the Chamaetia/Vetrix clade which includes subg. Chamaetia and Vetrix along with Salix arbutifolia (Chosenia), S. cardiophylla (Toisusu) and sect. Amygdalinae. Furthermore, interspecific hybridization occurs mostly within the Vetrix or Salix clades (Wagner et al., 2021a), raising the question whether the two main clades with different SDSs have reproductive barriers.

Thus, in the present study we aimed: (1) to use WGR data to reconstruct the phylogeny of sampled specimens (this approach has never been applied to the reconstruction of willow phylogeny across the whole genus); (2) to test whether phylogenies based on autosomes, SLRs and chloroplast sequences are consistent with one another; and (3) to determine whether differences in SDSs between species of the two clades offer a reliable character for their subgeneric division, or not.

MATERIALS AND METHODS

Taxon sampling

We included 90 Salix samples (23 sections, 62 species and eight varieties) in our analysis (Supplementary Data Table S2); 59 samples representing 48 taxa were newly collected from China, Japan and North America for this study. The species represent the five previously recognized subgenera Salix, Protitea, Vetrix, Longifoliae and Chamaetia. The plant material was frozen in liquid nitrogen and stored at −80 °C until total genomic DNA was extracted or dried with silica gel. To cover more taxa and sections of the genus, resequencing genome data were included from 26 samples of L. He et al. (in preparation) and Guo et al. (2021), representing 23 taxa. All these taxa were identified using relevant floras and taxonomic papers (Fang et al., 1999; Ohashi, 2006; Argus, 2010; He et al., 2014, 2015; Liu et al., 2016, 2020; He and Chen, 2017; He, 2018; Zeng and He, 2020). We also downloaded whole genome sequencing data of five samples: Salix brachista (SRR7341535), S. dunnii (SRR12893418), S. purpurea (SRR3927002), S. suchowensis (SRR10197854) and S. viminalis (ERR1558612) from NCBI, which have available assembled genome(s) (reviewed in He et al., 2021b), and chose Populus euphratica (I_233/ SRR13324572) as the outgroup.

Ploidy determination

The ploidy level of 49 individuals representing 42 taxa was measured by flow cytometry (FCM). Salix polyclona (2x = 2n = 38, L. He et al., in preparation) was used as an external standard for ploidy determination. The FCM protocol of Doležel et al. (2007) was used. About 20–50 mg of silica gel-dried leaf tissue was incubated for 80 min in 1 mL LB01 buffer, and then chopped with a razor blade. The cell culture was then collected by gentle pipetting and filtered through a 38-µm nylon mesh. Before analysis, the samples were stained with 80 μg mL−1 propidium iodide (PI) simultaneously with 80μg mL−1 RNase in an ice bath for 30 min. About 5000 nuclei were measured for each sample.

Ploidy levels were estimated using a MoFlo-XDP flow cytometer (Beckman Coulter, Inc., Indianapolis, IN, USA), and FloMax V2.0 (Sysmex Partec GmbH, Münster, Germany) was used to evaluate the histograms for each sample. The ploidy level was obtained based on the following equation: Sample ploidy (integer) = reference ploidy × mean position of the G1 sample peak/mean position of the G1 reference peak. A maximum coefficient of variation (CV) value of 8 was accepted for each sample peak (G0/G1 peak), to control the quality of ploidy-level measurements.

Sequencing, reads mapping and variant calling

Total genomic DNA for 59 samples was extracted from fresh leaves frozen in liquid nitrogen and stored at −80 °C or silica gel-dried leaves using the Qiagen DNeasy Plant Mini Kit (Qiagen, Valencia, CA, USA) following the manufacturer’s instructions. WGR using paired-end libraries was performed on an Illumina NovaSeq 6000 with 150-bp read length on each end by NovoGene (Beijing, China) and Majorbio (Shanghai, China). The high-quality genome assemblies of Salix dunnii (female, including chr07X, clade Salix) (He et al., 2021b) and S. purpurea v5.1 (female, including phased Chr15Z and Chr15W, clade Vetrix) (Zhou et al., 2020), and chloroplast genomes of S. dunnii (He et al., 2021b) and S. purpurea (GenBank: KP019639.1) were used as reference genomes for all 90 samples of Salix and the outgroup in read mapping and variant calling.

The sequenced reads of all samples were filtered and trimmed using fastp, and reads with length <60 bp were discarded (Chen et al., 2018). Then clean reads were aligned to the genomes and chloroplast genome sequences of Salix dunnii and S. purpurea using the BWA-MEM algorithm from BWA 0.7.12 (Li and Durbin, 2009; Li, 2013). SAMTOOLS 0.1.19 (Li et al., 2009) was used to extract primary alignments, and to sort and merge the mapped data. We used Sambamba 0.7.1 (Tarasov et al., 2015) to mark potential duplications from the PCR amplification step of library preparation for all bam files.

We called variants for all the bam files using GATK’s ‘HaplotypeCaller’ and ‘GenotypeGVCFs’ (Genome Analysis Toolkit v.4.1.8.1). For each bam file, ‘HaplotypeCaller’ for chromosome regions was run with ‘--sample-ploidy 2’, and for chloroplast regions with ‘--sample-ploidy 1’. Genomic VCFs (GVCFs) of chromosome and chloroplast regions were obtained for each sample. Before joint genotyping with ‘GenotypeGVCFs’, ‘GenomicsDBImport’ and ‘CombineGVCFs’ were used to merge the GVCFs of chromosome regions and chloroplast regions from all samples, respectively. Hard filtering of the single nucleotide polymorphisms (SNPs) was carried out using the best practice quality recommendations of the GATK group (QD < 2.0, FS > 60.0, MQ < 40.0, MQRankSum < −12.5, ReadPosRankSum < −8.0, SOR > 3.0). Biallelic sites were extracted for subsequent filtering for the chromosome regions. For the sites in the chloroplast regions, we kept all polymorphisms. We then excluded sites with extremely high coverage across all samples (twice the average coverage) and treated the sample-level genotype depth ‘<4’ as no call, and included sites with at most 10 % of no-call genotypes in all sample (for sites in chloroplast sequences, we allowed 50 % maximum missing). We also removed sites with a minor allele frequency <0.05. This process yielded four high-quality SNP datasets: 416 SNPs using the chloroplast genome of Salix dunnii as reference (CP-dun), 3 036 086 SNPs using the S. dunnii nuclear genome as the reference, 988 SNPs using the chloroplast genome of S. purpurea as reference (CP-pur) and 3 350 756 with the S. purpurea nuclear genome reference.

Phylogenetic analysis

Since different SDSs were identified in the two main diploid clades, we extracted the SNPs in autosomal regions from both nuclear datasets (excluding the entire chromosomes 7 and 15). A python script (https://github.com/zhangrengang/degeneracy) was used to generate all four-fold degenerate (FF-D) sites in the genomes of Salix dunnii and S. purpurea, based on their gene annotation (GFF3 files, Zhou et al., 2020; He et al., 2021b). We extracted the SNPs at these FF-D sites from the two nuclear autosomal datasets. Polyploidy can confuse phylogenetic analysis, especially if interspecific hybridization was involved in duplication of the genome, which can lead to incongruent reconstructions as well as reticulate evolutionary patterns (Alix et al., 2017). We therefore obtained 233 684 FF-D SNPs for all individuals, and 207 155 FF-D SNPs for diploids, from the autosome regions AR-dun (SNPs in autosome regions, using Salix dunnii as reference genome); similarly, we obtained 258 908 FF-D SNPs for all individuals and 230 084 for diploids from the AR-pur (SNPs in autosome regions, using S. pupurea as reference genome).

Phylogenetic relationships were inferred by a maximum-likelihood approach using RAXML v0.8.2.4 (Stamatakis, 2014) based on the concatenated sequences of six datasets: AR-dun (all individuals), AR-Di-dun (diploids only), CP-dun, AR-pur (all individuals), AR-Di-pur (diploids only) and CP-pur. Support values were calculated using 100 rapid bootstrap replicates (-f a option) based on the GTR + GAMMA nucleotide substitution model as used by Wagner et al. (2020) and He et al. (2021a) for the phylogenomics of willows.

The SDSs are not known for any polyploid willows. We therefore extracted SNPs from the SLRs only for the diploids. For the dataset using the S. dunnii reference, this region is X-LR, chr07:5675000–8880000 of chromosome 7, and for that using the S. purpurea reference it is the Z-linked region of chromosome 15 (Z-LR, Chr15Z: 2341099–6715814) (Zhou et al., 2020; He et al., 2021b). Because recombination suppression of SLRs may lead to gene duplications, we used only single-copy genes (SCGs) in the chromosome 7X and 15Z regions, identified using OrthoFinder (Emms and Kelly, 2019). Because of assembly quality and availability of phased SLRs, we used only the genomes of S. brachista, S. dunnii and S. purpurea to identify the chromosome 7X SCGs, and only S. dunnii and S. purpurea (excluding Chr15W) to identify the chromosome 15Z SCGs. Then we used the two SCG datasets to extract the SNPs in the X-LR and Z-LR. This yielded 1165 SNPs in 32 SCGs in X-LR, and 663 SNPs in 15 of 74 SCGs in Z-LR. We developed a custom script to obtain the sequences of each SCG in these datasets, and used ModelFinder (Kalyaanamoorthy et al., 2017) to select the best model under the Bayesian information criterion, and IQ-tree based on the best model (Minh et al., 2020) to construct individual gene trees. Species trees of the two datasets were estimated using ASTRAL (Zhang et al., 2018a).

RESULTS

Ploidy determination

Among the 42 measured taxa, the ploidy levels of 17 species and five varieties were measured for the first time (Supplementary Data Table S2, Fig. S1). Salix annulifera, S. baileyi, S. caroliniana, S. cheilophila, S. fargesii var. hypotricha, S. fargesii var. kansuensis, S. hypoleuca var. kansuensis, S. hypoleuca var. platyphylla, S. luctuosa, S. permollis, S. spathulifolia and S. tangii were revealed to be diploid. Salix austrotibetica, S. balansaei, S. chienii, S. paraplesia, S. sclerophylloides, S. spathulifolia var. glabra and S. staintoniana were found to be tetraploid. Salix taipaiensis, S. wangiana and S. yuhuangshanensis were hexaploid. Of the other 20 taxa, 17 had ploidy levels congruent with previous reports (Table S2). However, the ploidy levels of tetraploid Salix daltoniana, diploid S. cf. scouleriana and hexaploid S. sinica differed from those previously reported [S. daltoniana diploid (Fang et al., 1999); S. scouleriana tetraploid (Argus, 2010); and S. sinica tetraploid (L. He et al., in preparation)].

Whole genome re-sequencing

After filtering, we obtained more than 7 (mean 11) Gb of clean reads per new sequenced sample (Supplementary Data Table S3), and on average 11.4 Gb of all samples. The average depths of the new sequences ranged from 14.5× to 44.1× and from 11.9× to 44.2×, using Salix dunnii and S. purpurea as reference genomes, respectively, while all the 91 samples had average depths of 14.5– 62.9× and 11.9–63.5×, respectively (Tables S4 and S5). Based on more than 1 TB (0.65 TB newly sequenced) clean reads, we obtained 416 chloroplast and 3 036 086 nuclear high-quality SNPs using S. dunnii as reference, 988 and 3 350 756 high-quality SNPs with S. purpurea as reference.

Phylogenetic trees based on nuclear and chloroplast data

As we hypothesized that the species of the two main clades of Salix have different SDSs (see Introduction), we chose Salix dunnii and S. purpurea from the two clades as reference genomes in order to analyse SLRs separately from autosomal ones, and used a total of eight datasets to reconstruct phylogenies and species trees of Salix species (see Methods; Table 2, Figs 14; and Supplementary Data Figs S2–S5).

Table 2.

Details of the eight datasets used to conduct phylogenetic analyses

Dataset
Nuclear genome
Four-fold degenerate SNPs of autosomal region
 AR-dun233 684
 AR-pur258 908
Four-fold degenerate SNPs of autosomal region (diploids only)
 AR-Di-dun207 155
 AR-Di-pur230 084
SNPs of single-copy genes in sex-linked region (diploid only)
 SCG-X-LR1165
 SCG-Z-LR663
Chloroplast genome
 CP-dun416
 CP-pur988

Dun and X-LR used the Salix dunnii genome as reference; X-LR represents the X-linked region on chromosome 7 of Salix dunnii. Pur and Z-LR used the Salix purpurea genome as reference; Z-LR represents the Z-linked region on chromosome 15 of Salix purpurea. AR, CP and SCG represent autosomal region, chloroplast genome and single copy gene, respectively.

An external file that holds a picture, illustration, etc.
Object name is mcac012_fig1.jpg

Phylogeny inferred for 70 taxa and 91 samples of the genus Salix and the outgroup Populus euphratica based on maximum-likelihood analyses of the AR-dun (using the genome of Salix dunnii as reference, excluding chromosomes 7 and 15) dataset (233 684 four-fold degenerate SNPs) using RAxML. The subgenera are according to Skvortsov (1999), Ohashi (2006) and Argus (2010). Fang et al. (1999) divided Salix species only at the sectional level, but not all of those sections are monophyletic groups (He et al., 2021a). Thus, only species that were confirmed to belong to particular subgenera, based on Skvortsov (1999), Ohashi (2006) and Argus (2010), were coloured, accordingly. The scale bar refers the mean number of nucleotide substitution per site. Numbers above branches refer to bootstrap support values.

An external file that holds a picture, illustration, etc.
Object name is mcac012_fig4.jpg

Phylogeny inferred for 70 taxa and 91 samples of the genus Salix and the outgroup Populus euphratica based on maximum-likelihood analyses of the CP-dun (using the chloroplast genome of Salix dunnii as reference) dataset (416 SNPs) using RAxML. The scale bar refers the mean number of nucleotide substitution per site. Numbers above branches refer to bootstrap support values.

Five major clades and one paraphyletic group were found by analyses of putatively autosomal sequences, using both the AR-dun and AR-pur datasets, which include both diploid and polyploid species (Fig. 1; Supplementary Data Fig S2). Clade Vetrix includes four subclades: C1, C2, C3 and C4. Salix includes one subclade (C6) and a small polyploid group, C5, consisting of Salix shihtsuanensis, S. chienii and S. matsudana, plus three branches with S. cf. fragilis, S. pentandra, S. paraplesia and S. lucida, paraphyletic to the Vetrix clade. Individuals of the same species are well supported as monophyletic in the two trees, except a few species or complex with multiple ploidy levels (S. polyclona, S. shihtsuanensis), allopolyploid origin (S. opsimantha) or unclear morphological boundaries (S. luctuosa and S. hypoleuca; S. fargesii and S. moupinensis).

Compared to the complete autosome datasets, the data from just the diploid species [AR-Di-dun, AR-Di-pur, and SLR datasets (SCG-X-LR and SCG-Z-LR)] are expected to be less affected by the problems outlined above for polyploid species, such as paraphyly and reticulate evolution. All analyses of the diploid data support separate monophyletic groups of Vetrix and Salix species (Figs 2 and and3;3; Supplementary Data Figs S3 and S4). Due to lower numbers of SNPs, the SCG-X-LR and SCG-Z-LR species trees did not distinguish C1 and C2, but revealed C3 as sister to the C1 and C2 groups. The AR-Di-dun and AR-Di-pur analyses revealed a similar topology of C1, C2, C3, and C4 of clade Vetrix as with the whole autosome datasets.

An external file that holds a picture, illustration, etc.
Object name is mcac012_fig2.jpg

Phylogeny inferred for 38 diploid taxa and 49 samples of the genus Salix and the outgroup Populus euphratica based on maximum-likelihood analyses of the AR-Di-dun (using the genome of Salix dunnii as reference, excluding chromosomes 7 and 15) dataset (207 155 four-fold degenerate SNPs) using RAxML. The scale bar refers the mean number of nucleotide substitution per site. Numbers above branches refer to bootstrap support values.

An external file that holds a picture, illustration, etc.
Object name is mcac012_fig3.jpg

Phylogeny inferred for 38 taxa and 49 diploid samples of the genus Salix and the outgroup Populus euphratica based on the SCG-X-LR dataset (using the genome of Salix dunnii as reference, 32 single-copy genes, 1165 SNPs) analysed with ASTRAL species tree methods. Values above branches refer to local posterior probabilities.

Both trees based on chloroplast genomes, CP-dun and CP-pur, again included two major well-supported clades, Salix and Vetrix (Fig. 4; Supplementary Data Fig. S5). Bootstrap support values for most of the internal nodes within the major clades were low, and no subclades were clearly defined within Vetrix, with the notable exception of subclade P1, Salix triandra. Representatives of subg. Chamaetia and Vetrix were intermixed in one highly supported clade. In contrast, the Salix clade was fully supported as distinct; all species previously recognized as members of subg. Salix (including S. exigua and S. interior) fell within it, in two well-supported subclades.

Discussion

Phylogenomic analyses of Salix

Five major clades and one paraphyletic group were formed in autosome data-based trees (Fig. 1; Supplementary Data Fig. S2). Subclade C1 mainly comprises Asian species, especially endemic species of the Qinghai–Tibetan Plateau (QTP, including Hengduan Mountains and Himalaya), generally supporting the finding of a radiation by He et al. (2021a) in their Hengduan Mountains (HDM) clade, whereas C2 includes Eurasian and North American willows (Fig. 1; Fig. S2, Table S2). The species of both subclades are present in the Vetrix clade of chloroplast genome-based trees (Fig. 3; Fig. S4). The American species Salix exigua and S. interior (C3) have been considered to belong to subg. Longifoliae (Argus, 2010), but their position within Vetrix is highly supported in our analysis. This is incongruent with the chloroplast genome trees and may be caused by interspecific hybridization or incomplete lineage sorting of nuclear gene copies (see ‘Placement of section Longifoliae’). Salix triandra is the only species in C4, and appears as a sister to C3 and the other Vetrix representatives, which is consistent with previous studies (Chen et al., 2010; Hardig et al., 2010; Wu et al., 2015; Lauron-Moreau et al., 2015a; Wagner et al., 2021b; see also ‘Placement of Salix triandra’).

Members of subg. Salix as circumscribed by Argus (2010) formed the C5 grade, which is paraphyletic to the Vetrix clade. Hence, subg. Salix sensu Argus is a paraphyletic grade when the polyploid C5 group is included. In contrast, in the chloroplast genome trees the species of C5 were embedded within Salix, subclades P2 and P3 (Figs 1 and and4;4; Supplementary Data Figs S2 and S5). This C5 grade comprises representatives of sections Pentandrae, Salix and Salix shihtsuanensis (and its varieties). Salix shihtsuanensis was incorrectly placed in section Sieboldianae based on morphological characters (Fang et al., 1999). The position of C5 is similar to the ITS tree of Lauron-Moreau et al. (2015b, correction ‘fig. 3’), in which the clade mainly consisting of subg. Salix species formed a sister group to clades Longifoliae and Chamaetia/Vetrix. The sampled taxa of C5 (Fig. 1; Fig. S2, Table S2) were all identified as polyploids, which may have affected their positions in the two nuclear-based trees. Polyploidy is further a major cause of paraphyly because parental diploid species and polyploid derivatives coexist, whereas reciprocal monophyly establishes only after extinction of ancestors (Hörandl, 2007; Hörandl and Stuessy, 2010). Salix matsudana of C5 appeared ~4 Mya and was predicted to be allotetraploid (Zhang et al., 2020). Considering the conflict of the nuclear and chloroplast trees (Figs 1 and and4;4; Figs S2 and S5), it is possible that members of the C5 subclade arose from allopolyploid offspring of crosses between species from clades Salix and Vetrix; admixture analysis of nuclear data support the hypothesis of a hybrid origin (L. He, unpubl. res.). However, at present no final taxonomic conclusion can be drawn on the C5 group, and how sex is determined in polyploid systems also need to be clarified, as well as the SDSs of more species.

Although chloroplast genome-based analysis supports two major robust clades within Salix, nuclear-based alternative topologies suggest a more complex subgeneric evolution, as expected. Nevertheless, the four trees based on diploid species only supported both Vetrix and Salix as monophyletic (Figs 2 and and3;3; Supplementary Data Figs S3 and S4). Taken together, representatives of Chamaetia/Vetrix grouped together in all our trees, suggesting their close affinity and supporting their merging into one clade Vetrix, despite the uncertain placements of Salix triandra and section Longifoliae (see below). It is also of note that since willows are widely distributed across the Northern Hemisphere (Skvortsov, 1999), for further investigation it is important to extend the geographical range of sampling in order to obtain a phylogeny that would more profoundly reflect subgeneric relationships within the genus on a worldwide scale.

Placement of Salix triandra

Although SDSs of a few species have been identified so far, representatives of two major groups (Salix and Vetrix) already exhibit different heterogamety. Species in clade Salix have an XX/XY system, whereas species in clade Vetrix have a ZW/ZZ system, which may act as a barrier to gene exchange (Stöck et al., 2021).

Salix triandra formed a sister branch to the rest of the Vetrix clade in all trees. For the past decade S. triandra has attracted great interest from willow taxonomists. The affiliation of this species to subg. Salix was put in doubt by Trybush et al. (2008), in whose study it fell out with Salix and Vetrix and formed a third cluster with approximately equal genetic similarity to both subgenera. Chen et al. (2010) and Wu et al. (2015) distinguished Triandrae clade (= sect. Amygdalinae), including S. triandra, as sister to the whole Chamaetia/Vetrix clade. Salix triandra has a female heterogamety (ZW) SDS (Li et al., 2020), the same as all the tested species of Vetrix (Table 1). Moreover, hybrids of Salix triandra and S. viminalis can produce viable seeds; however, no fertile cross was recorded with species of subgenus. Salix (S. alba, S. pentandra) (Karp et al., 2011). Nevertheless, diploid Salix triandra can hybridize with tetraploid S. fragilis from subg. Salix, resulting in a triploid hybrid (S. × alopecuroides), which is often found in Europe, because the two parental species frequently co-occur along rivulets (Rechinger, 1957; Neumann and Polatschek, 1972; Neumann, 1981; Dobes et al., 1997; Wagner et al., 2021a). Interestingly, Neumann (1981) reported that these hybrids often have catkins with both male and female flowers, which highlights the complexity of sex determination in polyploid willows as a result of interspecific crosses. The homomorphic sex chromosomes of willows may not act as a complete reproductive barrier (He et al., 2021b; Stöck et al., 2021). However, the SDS of S. fragilis has not yet been characterized, and it is unknown how SDRs actually work in polyploids. The fertility and abundance of S. × alopecuroides are unknown, but it does not form populations (E. Hörandl, pers. obs.).

The separated phylogenetic position and the high genetic divergence of Salix triandra from the other Vetrix species speaks against an inclusion of the species in subg. Vetrix. However, its inclusion in subg. Salix also seems unwarranted. Although we assigned it under the Vetrix clade in this study according to its SDS, further studies with more samples of the ‘Amygdalinae’ clade are needed to decide on a final taxonomic placement.

Placement of section Longifoliae

Skvortsov (1968) treated Longifoliae as a section of subg. Salix. He was the first to propose raising it to a subgenus, but he did not classify species outside Eurasia. He also claimed that species from this section possess a hypodermis lacking chloroplast on both sides of the leaves, which is similar to Chosenia (S. arbutifolia), which, in turn, belongs to the Chamaetia/Vetrix clade (Chen et al., 2010; Hardig et al., 2010). However, Azuma et al. (2000) considered the possibility of independent evolution of this morphological feature in different lineages.

In hybridization experiments, Mosseler (1989, 1990) used a selection of species from the Salix and Vetrix clades, including S. exigua/interior (Argus, 2010). The latter exhibited pollen–pistil incongruity in crosses with members of clade Salix, resulting in seed abortion. In contrast, S. exigua/interior did not show a pollination barrier in crosses with diploid representatives of clade Vetrix (S. eriocephala and S. petiolaris), and produced viable F1 progeny. These studies suggest a closer affinity of S. exigua/interior with clade Vetrix than with Salix.

No previous molecular phylogenetic studies of Salix have proposed to treat Longifoliae as a part of Vetrix. Chong et al. (1995), examining allozyme variation in order to estimate genetic distance between S. exigua and other North American willows, found that the species was equally distant from both Vetrix and Salix, suggesting the revision of its taxonomical placement. In Leskinen and Alström-Rapaport (1999), S. exigua fell out of the main group of Salix species, so the authors suggested its earlier divergence. In the ITS tree of Lauron-Moreau et al. (2015b), four major clades were formed. One of them consisted of most Longifoliae species and appeared as a sister clade to Chamaetia/Vetrix, which, however, was not consistent with the plastid-based tree, in which Longifoliae perfectly grouped with members of clade Salix.

If we assume that differences in the SDSs lead to reproductive barriers between two clades (clade Salix and clade Vetrix), S. exigua/interior, which was found to inter-cross successfully with clade Vetrix species producing viable and vigorous F1 progeny (Mosseler, 1989; 1990), probably has a female heterogamety as do species in the Vetrix clade investigated so far. However, due to the lack of research on SDSs of any species of Longifoliae, this hypothesis needs to be tested. It is highly likely that phylogenetic incongruence between S. exigua and S. interior in chloroplast and nuclear trees appeared because of early interspecific hybridization or incomplete lineage sorting of nuclear gene copies (Figs 1 and and4;4; Supplementary Data Figs S2 and S5). Therefore, whether S. exigua and S. interior are the only two unique species of Longifoliae, or all species of this subgenus (section) have the same pattern of sex incapability when used in interspecific hybridizations, remains in question. Further identification of the SDSs will help to investigate our hypothesis, especially of species of sections Longifoliae and Amygdalinae.

Reproductive barriers via SDSs in willows

In recent years, both male and female heterogamety have been found within Salix (see Table 1). Based on our analyses, we hypothesize that the two major groups, Salix and Vetrix, might have different male and female heterogamety, respectively. One exception is Salix triandra (female heterogamety; Li et al., 2020), whose taxonomic position remains uncertain (see ‘Placement of Salix triandra’ above). We next discuss whether the sex-determination difference between the two major clades in Salix might act as a reproductive barrier.

Various mechanisms can limit plant hybridization, and both pre- and postzygotic barriers are known (Baack et al., 2015); and prezygotic reproductive barriers such as different elevation preferences (Wagner et al., 2021a) have been documented between willow species in the subgenera Vetrix and Salix.Mosseler and Papadopol (1989) concluded that common prezygotic barriers, such as ecological or spatial isolation, did not act, since a wide range of hybrids was found in natural populations within a 100-km area around Toronto. On the other hand, the same study recorded the flowering times of seven species from the subgenera Vetrix and Salix, and revealed two phenological groups: early-flowering Vetrix species and late-flowering willows belonging to Salix, suggesting that this could be a prezygotic barrier preventing interspecific hybridization between these two groups of species. Moreover, Mosseler (1989) demonstrated pollen–pistil incongruity between species of Vetrix and Salix, as successful pollinations were rare between representatives of these subgenera.

Given the impact of polyploidy on hybridization, in general, different ploidy levels produce strong crossing barriers in willows (Wagner et al., 2021a). On the other hand, homoploid crosses of species within the same subgenus usually yield viable seeds (Argus, 1974; Mosseler, 1990; Choudhary et al., 2013; Gramlich and Hörandl, 2016). For postzygotic barriers it is difficult to disentangle crossing barriers produced by different ploidy levels from those produced by putative different SDSs, and hence we focus here on crosses on the same ploidy level between the groups.

Data concerning postzygotic barriers and viability of seeds and hybrids obtained from interspecific crosses are scarce. Argus (1974) carried out pollinations between four Salix species. Among them, in crosses between tetraploid S. discolor from subg. Vetrix and S. lucida from subg. Salix (4x; Dorn, 1976), no seeds were produced. Diploid Salix exigua (sect. Longifoliae, subg. Salix, Argus, 2010) expressed neither pre- nor postzygotic barriers with diploid representatives of subg. Vetrix (Mosseler and Papadopol, 1989; Mosseler, 1989, 1990), suggesting that it may belong to this subgenus (which is consistent with the trees based on nuclear sequences in Figs 1–3 and Supplementary Data Figs S2S4, but not with the chloroplast result in Fig. 4 and Fig. S5). Salix humboldtiana, a diploid South American species from subg. Salix (Argus, 1997), not studied here, was used in breeding experiments for developing multipurpose willows. When used as a pistillate parent, it produced progeny, though weak, when pollinated by diploid S. viminalis, S. purpurea and S. daphnoides (also from the Vetrix clade; not studied here) (Argus, 2010; Bubner et al., 2018; Förster et al., 2021). Finally, the allopolyploid origin of the C5 grade in our phylogeny could have resulted from parental lineages of the diploid Vetrix and Salix clades.

Salix fits a single-factor model of the SDS (He et al., 2021b; Renner and Müller 2021; Sanderson et al., 2021). According to Renner and Müller (2021), females in the ZW system are heterozygous for a dominant W-linked femaleness factor. In the XY system, a Y-linked specific factor in the heterozygous males dominantly suppresses female functions. A hypothetical scheme in Supplementary Data Figure S6 shows the four offspring genotype classes when a zW female is inter-crossed with an xY male. Only two of them, with W (15, dominant) and x (7, recessive), and z (15, recessive) and Y (7, dominant) have no conflict between the dominance of the SD factors on the different chromosomes, and (assuming that presence of a chromosome 7 has no effect on sex in the species whose SD locus is on chromosome 15 unless it carries a dominant factor, and vice versa) might be expected to be female and male, respectively. The zx lacks the W femaleness-specific factor, but also lacks the Y-specific factor, so it could be female. Although WY were observed in fish (Xiphophorus maculatus) (Kallman 1984), its YW individuals can be female or male. However, SDSs in plants and animals evolved in different ways (Mank, 2022). The outcome cannot be predicted with certainty for the WY of willows, too. In the reciprocal cross (zz male × xx female), all offspring are xz. Hence, even if these hybrids were viable, fertility would be very low, especially in competition with the parental population. Even if fertile, there would probably be an excess of female genotypes. Eventually, polyploidy could overcome expression bias of the diploid hybrid by gene duplication, or two locus systems could evolve; such mechanisms could have helped the allopolyploid C5 group to establish and to evolve fertile species, but this needs to be studied.

The rate of sex chromosome divergence is neither unidirectional nor correlated with time (reviewed by Mank, 2022). In willows, homomorphism of sex chromosomes does not mean that they have recently evolved (He et al., 2021b; Mank, 2022). Although turnover events are possible within each clade of Salix (Renner and Müller, 2021), it is reasonable to assume that species within each clade share the same ancestral SDS on the same chromosome (Almeida et al., 2020). SDSs of willows may have contributed to isolation of the two major groups. A comprehensive dated tree of willows and divergence time estimation of X (Z) and Y (W) with a broader sampling should further aid our understanding of the correlation between SDSs and Salix diversification.

Conclusions

Overall, within the genus Salix there are the two clearly defined groups, Vetrix and Salix. The former comprises four subclades: endemic Asian species (C1); Eurasian and North American species (C2); two species of sect. Longifoliae (C3); and Salix triandra (C4). The Salix group becomes paraphyletic by inclusion of the mainly polyploid species (C5), and includes a group with mainly species of subg. Protitae (C6). Our analysis suggested that species expressing female or male heterogamety belong to different clades. However, it is not clear whether the difference in heterogamety is a barrier to hybridization. This uncertainty is partly due to the fact that the SDSs of only seven species have been identified, while the type of heterogamety remains unknown for many species used in breeding experiments, and partly because it remains unclear how ploidy affects interspecific mating (see ‘Placement of Salix triandra’). Furthermore, the C5 polyploids (Fig. 1; Supplementary Data Fig. S2, Table S2) could have originated from hybridization of species of Salix and Vetrix. If so, this would support incompleteness of the postzygotic reproductive barriers due to different SDSs between willows of the Salix and Vetrix clades.

SUPPLEMENTARY DATA

Supplementary data are available online at https://academic.oup.com/aob and consist of the following. Figure S1: Selected flow cytometry histograms of the estimated samples. Figure S2: Phylogeny inferred for 70 taxa and 91 samples of the genus Salix and the outgroup Populus euphratica based on maximum-likelihood analyses of the AR-pur dataset using RAxML. Figure S3: Phylogeny inferred for 38 diploid taxa and 49 samples of the genus Salix and the outgroup Populus euphratica based on maximum-likelihood analyses of the AR-Di-pur dataset using RAxML. Figure S4: Phylogeny inferred for 38 taxa and 49 diploid samples of the genus Salix and the outgroup Populus euphratica based on the SCG-Z-LR dataset analysed with ASTRAL species tree methods. Figure S5: Phylogeny inferred for 70 taxa and 91 samples of the genus Salix and the outgroup Populus euphratica based on maximum-likelihood analyses of the CP-pur dataset using RAxML. Figure S6: Hypothetical scheme of crosses between the ZW/ZZ and XX/XY systems. Table S1: Recent significant treatments of Salix with a focus on subgeneric classification. Table S2: Details of plant materials used in this study. Table S3: Statistics of quality control results of whole genome sequencing datasets of 91 samples. Table S4: Summary of mapping results of 91 samples using Salix dunnii as a reference genome. Table S5: Summary of mapping results of 91 samples using Salix purpurea as a reference genome.

mcac012_suppl_Supplementary_Figure_S1

mcac012_suppl_Supplementary_Figure_S2

mcac012_suppl_Supplementary_Figure_S3

mcac012_suppl_Supplementary_Figure_S4

mcac012_suppl_Supplementary_Figure_S5

mcac012_suppl_Supplementary_Figure_S6

mcac012_suppl_Supplementary_Table_S1

mcac012_suppl_Supplementary_Table_S2

mcac012_suppl_Supplementary_Table_S3

mcac012_suppl_Supplementary_Table_S4

mcac012_suppl_Supplementary_Table_S5

Acknowledgments

We thank Deborah Charlesworth for her useful comments on our paper. We thank Shuai Liao, Dong Liu, Robbie E. Hart, Sheng-Jie Fu, Fang He, Jing Ma, Bin Chen, Andrew Townesmith, Chun-Lan Lian, Wan-Cheng Hou, Dan-Ni Zhao, Gongga Expedition, PE Xizang Expedition, Hao Zhang, Jun Zhao, Si-Wen Zeng, Yi Lei, Ce Shang and Tian Liang for sampling. We are indebted to Ren-Gang Zhang for his useful python script (https://github.com/zhangrengang/degeneracy). The sequence data presented in this article can be downloaded from the NCBI database under BioProject accession PRJNA765215. The developed custom script is available at: https://github.com/oasiswho/vcfbox.

Contributor Information

Sergey Gulyaev, College of Forestry, Fujian Agriculture and Forestry University, Fuzhou, 350002, China.

Xin-Jie Cai, College of Forestry, Fujian Agriculture and Forestry University, Fuzhou, 350002, China.

Fei-Yi Guo, College of Ecology and Nature Conservation, Beijing Forestry University, Beijing, 100083, China.

Satoshi Kikuchi, Hokkaido Research Center, Forestry and Forest Products Research Institute, Forest Research and Management Organization. Hitsujigaoka 7, Toyohira, Sapporo City, Hokkaido, 062-8516, Japan.

Wendy L Applequist, William L. Brown Center, Missouri Botanical Garden, St. Louis, MO 63110, USA.

Zhi-Xiang Zhang, College of Ecology and Nature Conservation, Beijing Forestry University, Beijing, 100083, China.

Elvira Hörandl, Department of Systematics, Biodiversity and Evolution of Plants (with Herbarium), University of Goettingen, Göttingen, 37073, Germany.

Li He, College of Forestry, Fujian Agriculture and Forestry University, Fuzhou, 350002, China. Eastern China Conservation Centre for Wild Endangered Plant Resources, Shanghai Chenshan Botanical Garden, Shanghai 201602, China.

FUNDING

This study was financially supported by the National Natural Science Foundation of China (grant Nos. 31800466, 32171813) and the Natural Science Foundation of Fujian Province of China (grant No. 2018J01613).

CONFLICT OF INTEREST

The authors declare that they have no conflict of interest.

LITERATURE CITED

  • Abdollahzadeh A, Osaloo KS, Maassoumi AA.. 2011. Molecular phylogeny of the genus Salix (Salicaceae) with an emphasize to its species in Iran. The Iranian Journal of Botany 17: 244–253. [Google Scholar]
  • Alix K, Gerard PR, Schwarzacher T, Heslop-Harrison JSP.. 2017. Polyploidy and interspecific hybridization: partners for adaptation, speciation and evolution in plants. Annals of Botany 120: 183–194. [Europe PMC free article] [Abstract] [Google Scholar]
  • Almeida P, Proux-Wera E, Churcher A, et al.  2020. Genome assembly of the basket willow, Salix viminalis, reveals earliest stages of sex chromosome expansion. BMC Biology 18: 1–18. [Europe PMC free article] [Abstract] [Google Scholar]
  • Argus  GW. 1974. An experimental study of hybridization and pollination in Salix (willow). Canadian Journal of Botany 52: 1613–1619. [Google Scholar]
  • Argus  GW. 1997. Infrageneric classification of Salix (Salicaceae) in the New World. Systematic Botany Monographs 52: 1–121. [Google Scholar]
  • Argus  GW. 2010. Salix. In: Flora of North America Editorial Committee, eds. Flora of North America North of Mexico. New York: Oxford University Press, 7: 23–51. [Google Scholar]
  • Azuma T, Kajita T, Yokoyama J, Ohashi H.. 2000. Phylogenetic relationships of Salix (Salicaceae) based on rbcL sequence data. American Journal of Botany 87: 67–75. [Abstract] [Google Scholar]
  • Baack E, Melo MC, Rieseberg LH, Ortiz-Barrientos D.. 2015. The origins of reproductive isolation in plants. New Phytologist 207: 968–984. [Abstract] [Google Scholar]
  • Barkalov VY, Kozyrenko MM.. 2014. Phylogenetic analysis of the Far Eastern Salix (Salicaceae) based on sequence data from chloroplast DNA regions and ITS of nuclear ribosomal DNA. Journal of Plant Science and Conservation 3: 3–19. [Google Scholar]
  • Boucher LD, Manchester SR, Judd WS.. 2003. An extinct genus of Salicaceae based on twigs with attached flowers, fruits, and foliage from the Eocene Green River Formation of Utah and Colorado, USA. American Journal of Botany 90: 1389–1399. [Abstract] [Google Scholar]
  • Bubner B, Köhler A, Zaspel I, et al.  2018. Breeding of multipurpose willows on the basis of Salix daphnoides Vill., Salix purpurea L. and Salix viminalis L. Landbauforschung 68: 53–66. [Google Scholar]
  • van Buijtenen JP, Einspahr DW.. 1959. Note on the presence of sex chromosomes in Populus tremuloides. Botanical Gazette 121: 60–61. [Google Scholar]
  • Chase MW, Zmarzty S, Lledo MD, Wurdack KJ, Swensen SM, Fay MF.. 2002. When in doubt, put it in Flacourtiaceae: a molecular phylogenetic analysis based on plastid rbcL DNA sequences. Kew Bulletin 57: 141–181. [Google Scholar]
  • Chen J-H, Sun H, Wen J, Yang Y-P.. 2010. Molecular phylogeny of Salix L. (Salicaceae) inferred from three chloroplast datasets and its systematic implications. Taxon 59: 29–37. [Google Scholar]
  • Chen S, Zhou Y, Chen Y, Gu J.. 2018. fastp: an ultra-fast all-in-one FASTQ preprocessor. Bioinformatics 34: i884–i890. [Europe PMC free article] [Abstract] [Google Scholar]
  • Chong DKX, Zsuffa L, Aravanopoulos FA.. 1995. Genetic relationship between Salix exigua and other North American willows (Salix L.): Evidence from allozyme variation. Biochemical Systematics and Ecology 23: 767–771. [Google Scholar]
  • Choudhary P, Singh NB, Verma A, Sharma JP.. 2013. Crossability relationship among tree willows (Salix spp.) and molecular genetic variation among their progenies. Indian Journal of Genetics and Plant Breeding 73:302–309. [Google Scholar]
  • Cronk QC, Needham I, Rudall PJ.. 2015. Evolution of catkins: inflorescence morphology of selected Salicaceae in an evolutionary and developmental context. Front Plant Science 6: 1030. [Europe PMC free article] [Abstract] [Google Scholar]
  • Dai X-G, Hu Q-J, Cai Q-L, et al.  2014. The willow genome and divergent evolution from poplar after the common genome duplication. Cell Research 24: 1274–1277. [Europe PMC free article] [Abstract] [Google Scholar]
  • Daniell H, Lin C-S, Yu M, Chang W-J.. 2016. Chloroplast genomes: diversity, evolution, and applications in genetic engineering. Genome Biology 17: 1–29. [Europe PMC free article] [Abstract] [Google Scholar]
  • Dobes C, Hahn B, Morawetz W.. 1997. Chromosomenzahlen zur gefässpflanzen-flora Österreichs. Linzer Biologische Beiträge 29: 5–43. [Google Scholar]
  • Doležel J, Greilhuber J, Suda J.. 2007. Estimation of nuclear DNA content in plants using flow cytometry. Nature Protocols 2: 2233–2244. [Abstract] [Google Scholar]
  • Dorn  RD. 1976. A synopsis of American Salix. Canadian Journal of Botany 54: 2769–2789. [Google Scholar]
  • Emms D, Kelly S.. 2019. OrthoFinder2: fast and accurate phylogenomic orthology analysis from gene sequences. Genome Biology 20: 238. 10.1186/s13059-019-1832-y [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
  • Fang C-F, Zhao S-D, Skvortsov AK.. 1999. Salicaceae. In: Wu ZY., Raven PH., eds. Flora of China. Beijing and St. Louis: Science Press and Missouri Botanical Garden Press; 4: 139–274. [Google Scholar]
  • Förster N, Antoniadou K, Zander M, et al.  2021. Chemoprofiling as breeding tool for pharmaceutical use of Salix. Frontiers in Plant Science 12: 579820. [Europe PMC free article] [Abstract] [Google Scholar]
  • Guo F-Y, Liu K-J, Wang Y-C, Li E-Z, Zhan Z-F, Zhang Z-X.. 2021. Complete chloroplast genome sequence of Salix sinopurpurea (Salicaceae). Mitochondrial DNA Part B 6: 718–719. [Europe PMC free article] [Abstract] [Google Scholar]
  • Gramlich S, Hörandl E.. 2016. Fitness of natural willow hybrids in a pioneer mosaic hybrid zone. Ecology and Evolution 6: 7645–7655. [Europe PMC free article] [Abstract] [Google Scholar]
  • Hardig TM, Anttila CK, Brunsfeld SJ.. 2010. A phylogenetic analysis of Salix (Salicaceae) based on matK and ribosomal DNA sequence data. Journal of Botany 2010: 1–12. [Google Scholar]
  • He  L. 2018. Taxonomy and nomenclature of Salix baileyi, S. rehderiana, and S. disperma. Phytotaxa 349: 54–60. [Google Scholar]
  • He L, Chen S-P.. 2017. Identity of Salix chingiana (Salicaceae) from China. Phytotaxa 323: 97–100. [Google Scholar]
  • He L, Jia K-H, Zhang R-G, et al.  2021.  b. Chromosome-scale assembly of the genome of Salix dunnii reveals a male-heterogametic sex determination system on chromosome 7. Molecular Ecology Resources 21: 1966–1982. [Europe PMC free article] [Abstract] [Google Scholar]
  • He L, Liao S, Zhang Z-X.. 2014. Taxonomic revision of Salix opsimantha Schneid. and Salix oreinoma Schneid. Acta Botanica Boreali-Occidentalia Sinica 34: 1904–1908. [Google Scholar]
  • He L, Liao S, Zhang Z-X.. 2015. Nomenclature and taxonomy of Salix floccosa, S. opsimantha and S. austrotibetica. Phytotaxa 201: 158–164. [Google Scholar]
  • He L, Wagner ND, Hörandl E.. 2021a. Restriction-site associated DNA sequencing data reveal a radiation of willow species (Salix L., Salicaceae) in the Hengduan Mountains and adjacent areas. Journal of Systematics and Evolution 59: 44–57. [Google Scholar]
  • Hörandl  E. 2007. Neglecting evolution is bad taxonomy. Taxon 56: 1–5. [Google Scholar]
  • Hörandl E, Stuessy TF.. 2010. Paraphyletic groups as natural units of biological classification. Taxon 59: 1641–1653. [Google Scholar]
  • Hou J, Ye N, Zhang D-F, et al.  2015. Different autosomes evolved into sex chromosomes in the sister genera of Salix and Populus. Scientific Reports 5: 1–6. [Europe PMC free article] [Abstract] [Google Scholar]
  • Isebrands JG, Richardson J.. 2014. Poplars and willows: trees for society and the environment. Oxford: CABI. [Google Scholar]
  • Kallman  KD. 1984. A new look at sex determination in poeciliid fishes. In TurnerBJ, editor. , eds. Evolutionary genetics of fishes. New York: Plenum Publishing, 95–171. [Google Scholar]
  • Kalyaanamoorthy S, Minh BQ, Wong TK, Von Haeseler A, Jermiin LS.. 2017. ModelFinder: fast model selection for accurate phylogenetic estimates. Nature Methods 14: 587–589. [Europe PMC free article] [Abstract] [Google Scholar]
  • Karp A, Hanley SJ, Trybush SO, Macalpine W, Pei M, Shield I.. 2011. Genetic improvement of willow for bioenergy and biofuels. Journal of Integrative Plant Biology 53: 151–165. [Abstract] [Google Scholar]
  • Lauron-Moreau A, Pitre FE, Argus GW, Labrecque M, Brouillet L.. 2015a. Phylogenetic relationships of American willows (Salix L., Salicaceae). PLoS One 10: e0121965. [Europe PMC free article] [Abstract] [Google Scholar]
  • Lauron-Moreau A, Pitre FE, Argus GW, Labrecque M, Brouillet L.. 2015b. Correction: phylogenetic relationships of American willows (Salix L., Salicaceae). PLoS One 10: e0138963. [Europe PMC free article] [Abstract] [Google Scholar]
  • Leskinen E, Alström-Rapaport C.. 1999. Molecular phylogeny of Salicaceae and closely related Flacourtiaceae: evidence from 5.8 S, ITS 1 and ITS 2 of the rDNA. Plant Systematics and Evolution 215: 209–227. [Google Scholar]
  • Li  H. 2013. Aligning sequence reads, clone sequences and assembly contigs with BWA-MEM. arXiv preprint arXiv:1303.399 7 [Google Scholar]
  • Li H, Durbin R.. 2009. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics 25: 1754–1760. [Europe PMC free article] [Abstract] [Google Scholar]
  • Li H, Handsaker B, Wysoker A, et al.  2009. The sequence alignment/map format and SAMtools. Bioinformatics 25: 2078–2079. [Europe PMC free article] [Abstract] [Google Scholar]
  • Li M-M, Wang D-Y, Zhang L, et al.  2019. Intergeneric relationships within the family Salicaceae s.l. based on plastid phylogenomics. International Journal of Molecular Sciences 20: 3788. [Europe PMC free article] [Abstract] [Google Scholar]
  • Li W, Wu H-T, Li X-P, Chen Y-N, Yin T-M.. 2020. Fine mapping of the sex locus in Salix triandra confirms a consistent sex determination mechanism in genus Salix. Horticulture Research 7: 1–12. [Europe PMC free article] [Abstract] [Google Scholar]
  • Liu L-J, He L, Applequist WL.. 2020. Untangling two Chinese Salix species (Salicaceae) published by CK Schneider, with lectotypification of four names. Willdenowia 50: 159–163. [Google Scholar]
  • Liu L-J, He L, Zhang Z-X.. 2016. Two new records of Salix L. from China. Acta Botanica Boreali-Occidentalia Sinica 36: 426–428. [Google Scholar]
  • Ma Z-Y, Wen J, Ickert-Bond SM, Nie Z-L, Chen L-Q, Liu X-Q.. 2018. Phylogenomics, biogeography, and adaptive radiation of grapes. Molecular Phylogenetics and Evolution 129: 258–267. [Abstract] [Google Scholar]
  • Malmstrøm M, Matschiner M, Tørresen OK, Jakobsen KS, Jentoft S.. 2017. Whole genome sequencing data and de novo draft assemblies for 66 teleost species. Scientific Data 4: 1–13. [Europe PMC free article] [Abstract] [Google Scholar]
  • Mank  JE. 2022. Are plant and animal sex chromosomes really all that different? Philosophical Transactions of the Royal Society, B, in press. [Europe PMC free article] [Abstract] [Google Scholar]
  • Minh BQ, Schmidt HA, Chernomor O, et al.  2020. IQ-TREE 2: new models and efficient methods for phylogenetic inference in the genomic era. Molecular Biology and Evolution 37: 1530–1534. [Europe PMC free article] [Abstract] [Google Scholar]
  • Morey M, Fernandez-Marmiesse A, Castineiras D, Fraga JM, Couce ML, Cocho JA.. 2013. A glimpse into past, present, and future DNA sequencing. Molecular Genetics and Metabolism 110: 3–24. [Abstract] [Google Scholar]
  • Mosseler  A. 1989. Interspecific pollen–pistil incongruity in Salix. Canadian Journal of Forest Research 19: 1161–1168. [Google Scholar]
  • Mosseler  A. 1990. Hybrid performance and species crossability relationships in willows (Salix). Canadian Journal of Botany 68: 2329–2338. [Google Scholar]
  • Mosseler A, Papadopol CS.. 1989. Seasonal isolation as a reproductive barrier among sympatric Salix species. Canadian Journal of Botany 67: 2563–2570. [Google Scholar]
  • Neumann  A. 1981. Die mitteleuropäischen Salix-Arten. Mitteilungen der forstlichen Bundesversuchsanstalt Wien 134: 1–152. [Google Scholar]
  • Neumann A, Polatschek A.. 1972. Cytotaxonomischer beitrag zur gattung Salix. Annalen des Naturhistorischen Museums in Wien 76: 619–633. [Google Scholar]
  • Ohashi  H. 2006. Salicaceae. In: Boufford D, Ohba H.., eds. Flora of Japan. Tokyo: Kodansha, IIa: 7–25. [Google Scholar]
  • Olofsson JK, Cantera I, Van de Paer C, et al.  2019. Phylogenomics using low-depth whole genome sequencing: a case study with the olive tribe. Molecular Ecology Resources 19: 877–892. [Abstract] [Google Scholar]
  • Raubeson LA, Jansen RK.. 2005. Plant diversity and evolution: genotypic and phenotypic variation in higher plants. Chloroplast genomes of plants. In HenryRJ, editor. , eds. Chloroplast genomes of plants. Wallingford: CABI Publishing, 45–68. [Google Scholar]
  • Rechinger  KH. 1957. Salix. In HegiG, editor. , eds. Illustrierte Flora von Mitteleuropa, 2. München: Carl Hanser Verlag, 44–135. [Google Scholar]
  • Renner SS, Muller NA.. 2021. Plant sex chromosomes defy evolutionary models of expanding recombination suppression and genetic degeneration. Nature Plants 7: 392–402. [Abstract] [Google Scholar]
  • Sanderson BJ, Feng G, Hu N, et al.  2021. Sex determination through X-Y heterogamety in Salix nigra. Heredity (Edinb) 126: 630–639. [Europe PMC free article] [Abstract] [Google Scholar]
  • Shendure J, Balasubramanian S, Church GM, et al.  2017. DNA sequencing at 40: past, present and future. Nature 550: 345–353. [Abstract] [Google Scholar]
  • Skvortsov  AK. 1968. Willows of the U.S.S.R. A taxonomic and geographic revision. Moscow: Publishing Office ‘Nauka’. [Google Scholar]
  • Skvortsov  AK. 1999. Willows of Russia and adjacent countries. Joensuu: University of Joensuu. [Google Scholar]
  • Stamatakis  A. 2014. RAxML Version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30: 1312–1313. [Europe PMC free article] [Abstract] [Google Scholar]
  • Stöck M, Dedukh D, Reifová R, Lamatsch DK, Starostová Z, Janko K.. 2021. Sex chromosomes in meiotic, hemiclonal, clonal and polyploid hybrid vertebrates: along the ‘extended speciation continuumʼ. Philosophical Transactions of the Royal Society B 376: 20200103. [Europe PMC free article] [Abstract] [Google Scholar]
  • Stull GW, Soltis PS, Soltis DE, Gitzendanner MA, Smith SA.. 2020. Nuclear phylogenomic analyses of asterids conflict with plastome trees and support novel relationships among major lineages. American Journal of Botany 107: 790–805. [Abstract] [Google Scholar]
  • Tarasov A, Vilella AJ, Cuppen E, Nijman IJ, Prins P.. 2015. Sambamba: fast processing of NGS alignment formats. Bioinformatics 31: 2032–2034. [Europe PMC free article] [Abstract] [Google Scholar]
  • Trybush S, Jahodová S, Macalpine W, Karp A.. 2008. A genetic study of a Salix germplasm resource reveals new insights into relationships among subgenera, sections and species. Bioenergy Research 1: 67–79. [Google Scholar]
  • Unamba CI, Nag A, Sharma RK.. 2015. Next generation sequencing technologies: the doorway to the unexplored genomics of non-model plants. Frontiers in Plant Science 6: 1074. [Europe PMC free article] [Abstract] [Google Scholar]
  • Wagner ND, Gramlich S, Hörandl E.. 2018. RAD sequencing resolved phylogenetic relationships in European shrub willows (Salix L. subg. Chamaetia and subg. Vetrix) and revealed multiple evolution of dwarf shrubs. Ecology and Evolution 8: 8243–8255. [Europe PMC free article] [Abstract] [Google Scholar]
  • Wagner ND, He L, Hörandl E.. 2020. Phylogenomic relationships and evolution of polyploid Salix species revealed by RAD sequencing data. Frontiers in Plant Science 11: 1077. [Europe PMC free article] [Abstract] [Google Scholar]
  • Wagner ND, He L, Hörandl E.. 2021a. The evolutionary history, diversity, and ecology of willows (Salix L.) in the European Alps. Diversity 13: 146–161. [Google Scholar]
  • Wagner ND, Volf M, Hörandl E.. 2021b. Highly diverse shrub willows (Salix L.) share highly similar plastomes. Frontiers in Plant Science 12: 662715. [Europe PMC free article] [Abstract] [Google Scholar]
  • Wu J, Nyman T, Wang DC, Argus GW, Yang YP, Chen JH.. 2015. Phylogeny of Salix subgenus Salix s.l. (Salicaceae): delimitation, biogeography, and reticulate evolution. BMC Evolutionary Biology 15: 1–13. [Europe PMC free article] [Abstract] [Google Scholar]
  • Zeng S-W, He L.. 2020. Lectotypification and morphological notes of Salix annulifera C. Marquand & Airy Shaw (Salicaceae). Journal of Tropical and Subtropical Botany 28: 236–240. [Google Scholar]
  • Zhang C, Rabiee M, Sayyari E, Mirarab S.. 2018a. ASTRAL-III: polynomial time species tree reconstruction from partially resolved gene trees. BMC Bioinformatics 19: 15–30. [Europe PMC free article] [Abstract] [Google Scholar]
  • Zhang L, Xi Z, Wang M, Guo X, Ma T.. 2018b. Plastome phylogeny and lineage diversification of Salicaceae with focus on poplars and willows. Ecology and Evolution 8: 7817–7823. [Europe PMC free article] [Abstract] [Google Scholar]
  • Zhang J, Yuan H, Li Y, et al.  2020. Genome sequencing and phylogenetic analysis of allotetraploid Salix matsudana Koidz. Horticulture Research 7: 201. [Europe PMC free article] [Abstract] [Google Scholar]
  • Zhang Q, Liu Y.. 2003. Examination of the cytoplasmic DNA in male reproductive cells to determine the potential for cytoplasmic inheritance in 295 angiosperm species. Plant and Cell Physiology 44: 941–951. [Abstract] [Google Scholar]
  • Zhou R, Macaya-Sanz D, Carlson CH, et al.  2020. A willow sex chromosome reveals convergent evolution of complex palindromic repeats. Genome Biology 21: 1–9. [Europe PMC free article] [Abstract] [Google Scholar]

Articles from Annals of Botany are provided here courtesy of Oxford University Press

Citations & impact 


Impact metrics

Jump to Citations

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/122705325
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/122705325

Article citations


Go to all (7) article citations

Data 


Data behind the article

This data has been text mined from the article, or deposited into data resources.

Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.


Funding 


Funders who supported this work.

National Natural Science Foundation of China (2)

Natural Science Foundation of Fujian Province (1)