Rapid evolution of T2/S-RNase genes in Fragaria linked to multiple transitions from self-incompatibility to self-compatibility
Wu Chena,b, Hong Wanc, Fang Liua, Haiyuan Dua,b, Chengjun Zhanga,***, Weishu Fana,*, Andan Zhua,**     
a. Germplasm Bank of Wild Species, Kunming Institute of Botany, Chinese Academy of Sciences, Kunming, Yunnan 650201, China;
b. University of Chinese Academy of Sciences, Beijing 100049, China;
c. Horticultural Research Institute, Yunnan Academy of Agricultural Sciences, Kunming, Yunnan 650205, China
Abstract: The T2/RNase gene family is widespread in eukaryotes, and particular members of this family play critical roles in the gametophytic self-incompatibility (GSI) system in plants. Wild diploid strawberry (Fragaria) species have diversified their sexual systems via self-incompatible and self-compatible traits, yet how these traits evolved in Fragaria remains elusive. By integrating the published and de novo assembled genomes and the newly generated RNA-seq data, members of the RNase T2 gene family were systematically identified in six Fragaria species, including three self-incompatible species (Fragaria nipponica, Fragaria nubicola, and Fragaria viridis) and three self-compatible species (Fragaria nilgerrensis, Fragaria vesca, and Fragaria iinumae). In total, 115 RNase T2 genes were identified in the six Fragaria genomes and can be classified into three classes (Ⅰ–Ⅲ) according to phylogenetic analysis. The identified RNase T2 genes could be divided into 22 homologous gene sets according to amino acid sequence similarity and phylogenetic and syntenic relationships. We found that extensive gene loss and pseudogenization coupled with small-scale duplications mainly accounted for variations in the RNase T2 gene numbers in Fragaria. Multiple copies of homologous genes were mainly generated from tandem and segmental duplication events. Furthermore, we newly identified five S-RNase genes in three self-incompatible Fragaria genomes, including two in F. nipponica, two in F. viridis, and one in F. nubicola, which fit for typical features of a pistil determinant, including highly pistil-specific expression, highly polymorphic proteins and alkaline isoelectric point (pI), while no S-RNase genes were found in all three self-compatible Fragaria species. Surprisingly, these T2/S-RNase genes contain at least one large intron (> 10 kb). This study revealed that the rapid evolution of T2/S-RNase genes within the Fragaria genus could be associated with its sexual mode, and repeated evolution of the self-compatible traits in Fragaria was convergent via losses of S-RNase.
Keywords: Fragaria    RNase T2    Self-(in)compatibility    S-RNase    
1. Introduction

The RNase T2 gene family is a group of the ubiquitous enzyme RNases distributed broadly in animals, plants, protozoans, viruses, and some bacteria that catalyzes the cleavage of phosphodiester bonds in RNA to 3′ mononucleotides via 2′, 3′ cyclic nucleotides (Taylor et al., 1993; Irie, 1999; Deshpande and Shankar, 2002). RNase T2 genes participate in vital functions such as nutrition intake, phosphate remobilization, self-incompatibility (SI), senescence, and defense against pathogens (Irie, 1999; Deshpande and Shankar, 2002; Luhtala and Parker, 2010).

Members of the RNase T2 family in plants are classified into S-RNases and S-like RNases based on whether they are involved in gametophytic self-incompatibility (GSI) responses (Bariola and Green, 1997; MacIntosh, 2011). Phylogenetically, they can be categorized into three distinct classes: the T2/S-like RNases belong to classes Ⅰ and Ⅱ, while T2/S-RNases are in class Ⅲ (Igic and Kohn, 2001; Steinbachs and Holsinger, 2002; Roalson and McCubbin, 2003; MacIntosh et al., 2010; MacIntosh, 2011; Ramanauskas and Igic, 2017). S-like RNases show variable functions in different species. For example, they have been suggested to participate in the defense response in Nicotiana tabacum, while in tomato and Arabidopsis, they play a role in phosphate remobilization and nucleic acid scavenging (Jost et al., 1991; Köck et al., 1998; Chen et al., 2000; Kurata et al., 2002). S-RNase genes are the pistil determination components involved in S-RNase-based GSI, which could prevent self-fertilization (McClure et al., 1989; Sassa et al., 1996; Xue et al., 1996). The typical S-RNase genes are specifically and highly expressed in the pistil and encode alkaline, polymorphic extracellular ribonuclease. They have been characterized in various families, such as Solanaceae (McClure et al., 1989), Scrophulariaceae (Xue et al., 1996), Rosaceae (Sassa et al., 1993; Ishimizu et al., 1998), Rutaceae (Liang et al., 2020), and Cactaceae (Ramanauskas and Igić, 2021), etc. In Rosaceae, two different systems associated with GSI have been investigated: one is the self-recognition system in Prunus (e.g. almond, apricot, and cherry) and the other is the non-self-recognition system (e.g. apple and pear) (Sonneveld et al., 2005; Kubo et al., 2010; Greco et al., 2012; Ramanauskas and Igić, 2021). The S-RNase gene structure of the non-self-recognition system has one intron, while that of the self-recognition system has two introns (Sonneveld et al., 2003; Ortega et al., 2005; Mota et al., 2007; Dreesen et al., 2010). Although much attention has been given to the function of the RNase T2 family in Rosaceae, the genomic basis and evolutionary mechanisms of the T2/S-RNase gene family have been less studied. A recent study among seven Rosaceae species showed that the quantity of the RNase T2 gene in the self-incompatible species Prunus avium is twice than that in the self-compatible species Prunus persica (Zhu et al., 2020). This disparity implies that the number of the RNase T2 genes in self-incompatible species is probably higher than that in self-compatible species in a single genus, but more evidence is needed to confirm this hypothesis on a short phylogenomic scale.

Fragaria, commonly known as strawberries, is a genus that belongs to the Rosaceae family. The Fragaria genus comprises approximately 25 wild species including 12 diploid species (2n = 14) (Folta and Davis, 2006; Hummer and Hancock, 2009; Lei et al., 2016). All diploid Fragaria species are hermaphroditic and either self-incompatible or self-compatible species (Evans and Jones, 1967; Njuguna et al., 2013; Liston et al., 2014). The Fragaria genus engages in the S-RNase-based GSI system, and it has been speculated to be controlled by two independent loci (Boskovic et al., 2010; Du et al., 2019). However, in recent studies, two allelic S-RNases (Sa-RNase and Sb-RNase) have been identified, indicating that only one S locus exists in Fragaria viridis Duch. (Du et al., 2021). To address the evolutionary history of S-RNases, other S-RNase genes in Fragaria need to be identified. The released whole genome sequences of Fragaria vesca Lindl. (Edger et al., 2018; Li et al., 2019), Fragaria iinumae Makino. (Feng et al., 2021), Fragaria nilgerrensis Schlect. (Zhang et al., 2020), and Fragaria nubicola Lindl. (Feng et al., 2021) provide an opportunity to conduct the genome-wide identification of RNase T2 genes and to characterize the S-RNase genes involved in the self-incompatible response in a phylogenomic framework.

In this study, six Fragaria species including three self-incompatible species (Fragaria nipponica Lindl., F. nubicola, F. viridis) (Feng et al., 2021) and three self-compatible species (F. iinumae, F. nilgerrensis, F. vesca) (Edger et al., 2018; Li et al., 2019; Zhang et al., 2020; Feng et al., 2021) were selected to examine the evolution of RNase T2 genes at the whole genome level and to identify S-RNase genes involved in SI. By analyzing the phylogenetic relationship, physicochemical features, conserved motifs, duplication modes, and the expression of RNase T2 genes, we identified the S-RNase genes that are likely associated with SI and revealed the mechanisms underlying the rapid evolution of these genes.

2. Methods 2.1. Genomic data collection

The genome sequence and gene annotation of Fragaria nilgerrensis (version: WPAB01000000) (Zhang et al., 2020) was downloaded from the Genome Warehouse in the National Genomics Data Center (NGDC), Beijing Institute of Genomics (BIG), Chinese Academy of Sciences. Genome sequences of F. vesca (version: v4.0.a2) (Edger et al., 2018; Li et al., 2019), F. nubicola (version: v1.0) (Feng et al., 2021), F. iinumae (version: v1.0) (Feng et al., 2021), Potentilla micrantha (version: v1.0) (Buti et al., 2018), and Rosa chinensis (version: v1.0) (Saint-Oyant et al., 2018) were obtained from the Genome Database for Rosaceae (GDR, http://www.Rosaceae.org/). Genome sequences of F. viridis (unpublished) and F. nipponica (unpublished) were de novo assembled.

2.2. Identification and re-annotation of RNase T2 genes

A seed alignment file of the ribonuclease T2 domain (PF00445) was downloaded from the Pfam v35.0 database (http://pfam.janelia.org/) (Bateman et al., 2004) and was used to identify annotated RNase T2 proteins in the six Fragaria species and the two outgroups (R. chinensis and P. micrantha) through HMMER v3.3.1 (e-value < e−10) (Finn et al., 2011). Then, the Fragaria-specific RNase T2 HMM file was constructed using hmmbuild from the HMMER v3.3.1 based on the identified RNase T2 protein sequences, and a second-round identification of RNase T2 genes in the genomes was performed using HMMER v3.3.1 (e-value < e−10) and confirmed by HMMSCAN and the Conserved Domains Database (CDD v3.19) (https://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi) (Lu et al., 2020).

The identified RNase T2 genes were re-annotated using our unpublished RNA sequencing (RNA-seq) data of four Fragaria species (F. nilgerrensis, F. vesca, F. viridis and F. nipponica). For those RNase T2 genes without available RNA-seq data, a homology-based method was applied for re-annotation. The pseudogene identification was defined as more than one-third truncation relative to the intact cognate CDS, while the other identified genes were defined as functional genes. The deleterious mutation pattern resulting in pseudogenes was verified by next-generation resequencing data.

2.3. Phylogenetic analysis

A phylogeny was inferred from 5570 orthologs, which were identified by Orthofinder v2.5.2 (Emms and Kelly, 2015) with default parameters. The sequences were aligned by MAFFT v7.487 (Katoh, 2002) and Gblocks v0.91b (Castresana, 2000) was used to trim the poorly aligned regions with default parameters. The phylogenetic tree was constructed with the maximum likelihood (ML) method by PhyML v3.1 (Guindon et al., 2010) with 1000 bootstrap values.

The phylogeny of the RNase T2 gene family was generated as follows: complete coding nucleotide sequences of RNase T2 gene family members in the six Fragaria species were aligned by ClustalW (Thompson et al., 2003) with default parameters. The phylogenetic tree was constructed with the ML method implemented in MEGA 7.0.26 (Kumar et al., 2016) with 1000 bootstrap values.

2.4. Sequence features and analysis of protein properties

The intron-exon structures were plotted using the Gene Structure Display Server (GSDS 2.0, http://www.gsds.cbi.pku.edu.cn/) based on pairwise alignment of the genomic sequences and the CDS sequences (Hu et al., 2015). The conserved motifs from the full-length amino acid sequences of RNase T2 in the six Fragaria species were identified by the Multiple Expectation-Maximization for Motif Elicitation (MEME v5.0.5) tool (Bailey et al., 2009) with the following parameters: maximum number of different motifs, 15; minimum motif width, 6; and maximum motif width, 50. The results were plotted with TBtools v1.098696 (Chen et al., 2020). The molecular weight (MW) and isoelectric point (pI) were calculated using the ProtParam tool (https://web.expasy.org/protparam/) implemented in ExPASy (Gasteiger et al., 2005).

2.5. Gene location and synteny analysis

All RNase T2 genes of the six Fragaria species were mapped to their corresponding chromosomes, and their physical locations were plotted using MapGene2Chrom v2.0 (http://mg2c.iask.in/mg2c_v2.0/) (Jiangtao et al., 2015). The Multiple Collinearity Scan toolkit (MCScanX) (Wang et al., 2012) was adopted to analyze syntenic relationships and the gene duplication events of RNase T2 genes among the six Fragaria genomes with the default parameters.

2.6. RNA isolation, sequencing and gene expression analysis

Plants of Fragaria viridis and F. nipponica were grown in a greenhouse at Kunming Institute of Botany. Total RNA was separately isolated from styles, anthers, and mixed tissues (calyx and petals) at the balloon stage (Hollender et al., 2012) from the two selected Fragaria species by using an RNAprep Pure Plant Kit (polysaccharide- and polyphenolic-rich) (Tiangen, Beijing). Three biological replicates were generated for each tissue. The isolated RNA was used for high-throughput RNA-seq library construction, and sequenced (paired-end 150 bp) on an Illumina HiSeq 2500 platform. Clean data were de novo assembled using Trinity v2.9.1 (Grabherr et al., 2011) with default parameters, and were mapped to the genome using HISAT2 v2.0.5 (Kim et al., 2015). Gene expression values (TPM: transcripts per million) of the RNase T2 genes were calculated using StringTie v2.1.4 (Pertea et al., 2015).

2.7. Reverse transcription PCR analysis

To verify the tissue-specific expression of the identified S-RNase genes, multiple tissues from F. nipponica and F. viridis were collected, including roots, stems, leaves, calyxes, petals, anthers and pistils. Reverse transcription PCR (RT–PCR) analysis was performed using the PrimeScript One Step RT–PCR Kit (Takara, Dalian, China). The Actin gene of the two species was selected as the control. RT-PCR primers are collected in Table S6.

3. Results 3.1. Phylogenetic relationship of six diploid species suggests multiple transitions from self-incompatibility to self-compatibility in Fragaria

To investigate the origin of the self-compatible and self-incompatible species within Fragaria, three self-compatible species (F. nilgerrensis, F. vesca and F. iinumae) and three self-incompatible species (F. nipponica, F. nubicola and F. viridis) were selected for phylogenetic analysis, with R. chinensis and P. micrantha used as outgroups. A maximum likelihood (ML) phylogenetic tree was constructed using the concatenated dataset of 5570 single-copy genes (Fig. 1A). This resulted in a non-monophyletic group of self-incompatible species, showing that the self-incompatible species F. viridis is a sister lineage to the self-compatible species F. vesca with maximal bootstrap support (100%), and the self-compatible species F. nilgerrensis is sister to the two self-incompatible species (F. nipponica and F. nubicola). Thus, the self-compatible traits likely evolve independently in Fragaria.

Fig. 1 Phylogenetic relationship of six Fragaria species and the number of RNase T2 genes identified in each species. (A) An ML phylogenetic tree (on the left) of the selected Fragaria species inferred from 5570 single copy genes. The bootstrap support values were indicated on each node. The SC and SI species are highlighted with green and purple, respectively. Numbers of the identified RNase T2 genes (on the right) in the six Fragaria genomes are shown. (B) Synteny analysis of the 115 RNase T2 genes among the six Fragaria genomes. The species were ordered according to the inferred evolutionary relationship shown in Fig. A. The red triangles indicate the location of RNase T2 genes, and the blue lines link orthologous RNase T2 gene sets. The chromosome number is shown at the top of each Fragaria chromosome and ordered according to the collinearity to F. vesca. (C) An ML phylogenetic tree of the identified 115 RNase T2 genes. The bootstrap support values higher than 50 were labeled on each node. Class Ⅰ, Ⅱ, and Ⅲ are denoted by pink, blue and green, respectively. The RNase T2 gene number in each class is shown in parentheses. The gene names were named according to their species, the order on chromosomes and the order of homologous gene sets. The prefixes "Fiin", "Fnil", "Fnip", "Fnub", "Fves" and "Fvir" corresponded to F. iinumae, F. nilgerrensis, F. nipponica, F. nubicola, F. vesca and F. viridis, respectively. Genes with RNA data supported are signed with solid circles, while genes without RNA data supported are signed with solid triangles. Functional genes are colored by green, and pseudogenes are colored gray.
3.2. Identification and classification of RNase T2 genes in Fragaria genomes

A total of 115 RNase T2 genes were identified in the six Fragaria genomes including 28 homologous pseudogenes (Tables S1 and S2). Pseudogenized RNase T2 genes were mainly caused by frameshift mutations (67.86%), the gain of stop codons (14.28%), and exon loss (17.86%) (Table S2). We identified 27 RNase T2 genes (21 functional genes and six pseudogenes) in the self-incompatible species F. nipponica, ranking as the highest gene number. In contrast, only 14 RNase T2 genes were identified in the self-compatible species F. iinumae (10 functional genes and four pseudogenes). The total number of identified RNase T2 genes in F. nubicola, F. nilgerrensis, F. viridis and F. vesca was similar, with 18, 21, and 17 RNase T2 genes, respectively (Fig. 1A). The results revealed that the number of RNase T2 genes varied among Fragaria species.

A total of 22 homologous gene sets (HG1-22) were identified according to protein sequence similarity and phylogenetic relationships. These homologs were further verified by collinearity relationship (Fig. 1B). A phylogenetic tree inferred from all the 115 identified RNase T2 genes showed that they were separated into three distinct classes (Ⅰ–Ⅲ), among which class Ⅰ comprised the largest gene set (71 genes) distributed in 12 HGs (HG1-12) including 54 functional genes (presenting 76.06%), and 17 pseudogenes (presenting 23.94%); classes Ⅱ and Ⅲ contained 24 and 20 RNase T2 genes, respectively. Since T2/S-RNase genes are known to belong to Class Ⅲ (Ramanauskas and Igic, 2017; Du et al., 2021; Ramanauskas and Igić, 2021), we identified 15 functional RNase T2 genes in this class as the candidate S-RNase genes.

3.3. Motif composition, functional conserved domains and gene structure of RNase T2 genes in Fragaria genomes

The functional motifs of RNase T2 proteins were investigated, and a total of 15 conserved motifs were identified (Fig. S1). Class Ⅰ possesses 14 motifs including six unique motifs (motif 3, 8, 9, 11, 14, 15) while class Ⅱ and class Ⅲ showed highly similar motif compositions with nine and eight motifs, respectively. Motif 10 existed in class Ⅱ and class Ⅲ but not in class Ⅰ. Motif 2 and motif 1 corresponded to the two functional conserved domains (CAS Ⅰ and Ⅱ) of the RNase T2 protein, respectively. Multiple sequence alignments of the full-length amino acid sequences of 115 RNase T2 proteins indicated that most of these genes (71.3%) contained two acting histidine residues (Fig. S1B), and 33 (28.7%) members including 24 pseudogenes, lost at least one functionally conserved domain, probably suggesting the loss of gene function. Specifically, 11 RNase T2 proteins including eight pseudogenes lost both CAS Ⅰ and CAS Ⅱ (six in class Ⅰ, four in class Ⅱ, one in class Ⅲ). Another two subsets of genes have lost either the CAS Ⅰ or CAS Ⅱ domain. Five RNase T2 proteins including four pseudogenes (four in class Ⅰ, one in class Ⅲ), lost the CAS Ⅰ domain, and 17 RNase T2 proteins, including 12 pseudogenes, lost the CAS Ⅱ domain (eleven in class Ⅰ, four in class Ⅱ, two in class Ⅲ) (Fig. S1B). The loss of the CAS domain in RNase T2 proteins may cause partial or complete loss of their biological function.

Meanwhile, the intron-exon structure of RNase T2 genes showed substantial variations (Fig. S2). We found that the intact genes clustered into class Ⅰ (HG1-12) harboring two to eight exons, among which HG1-4 and HG6-7 possess two exons, HG8 and HG10 possess three exons, HG5, HG9 and HG11 possess four exons and HG12 possesses eight exons. For those genes that were clustered into class Ⅱ (HG13-18) and class Ⅲ (HG19-22), two to three exons were commonly identified. Specifically, HG13-20 possesses two exons, and HG21-HG22 possesses three exons (Fig. S2A and S2B). Exon loss may also affect the gene structures. For example, only one exon was found in FvirRNS9-HG1, a homologous pseudogene that belongs to HG1, among which intact genes had two exons (Fig. S2A).

3.4. Sequence features of RNase T2 genes in the Fragaria genomes

Genic characteristics, including the length of the protein sequence, molecular weight (MW) and isoelectric point (pI), of RNase T2 genes were analyzed (Table S1 and Fig. S3). The length of the 115 RNase T2 protein sequences ranged from 47 to 278 amino acids (Table S1). Previous studies showed that the majority of MWs of a typical RNase T2 enzyme are in the range of 20–40 kDa (Deshpande and Shankar, 2002). In this study, among the 115 identified RNase T2 proteins, 33 proteins were out of this range and displayed lower MWs including all 28 pseudogenes (Table S1). Moreover, the sequence lengths of the 33 low MW RNase T2 proteins were shorter than those of the rest of the RNase T2 proteins, which is consistent with their lower MWs. Outrageous values occurred in both self-incompatible and self-compatible species (Table S1), indicating irrelevant compatibility. The pI values ranged from 4.23 to 9.86 (Table S1 and Fig. S3). Class Ⅰ and Ⅱ proteins generally have a slightly higher proportion of alkaline pI values than acidic pI values; 38 (53.52%) and 13 (54.17%) peptides exhibit alkaline pI values, while 33 (46.48%) and 11 (45.83%) peptides exhibit acidic pI values, respectively. However, in class Ⅲ peptides, there were significantly more alkaline pI values; 16 (80%) peptides had alkaline pI values, while 4 (20%) peptides had acidic pI values. We found that only two homologous pseudogenes (FvirRNS9_HG1 and FnipRNS6_HG20) showed independent pI shifts from alkaline to acidic values, revealing that deleterious mutations have little effect on pI values. Our results also suggested that pI values were not associated with the loss of the active histidine residue (pseudogene).

3.5. Evolutionary mechanisms of copy number variations of RNase T2 genes in Fragaria genomes

We systematically examined the physical positions of the identified RNase T2 genes and found that they were unevenly distributed on the seven Fragaria chromosomes (Figs. 1B and S4). Taking F. nipponica as an example, for its total number of 27 RNase T2 genes, 22 genes (81.48%) were located on chr3, chr5, and chr7, and no RNase T2 genes were on chr1 and chr4 (Fig. S4). This uneven distribution of RNase T2 genes is consistent in all five other examined Fragaria genomes. Specifically, the colinear chromosomes to chr3, chr5 and chr7 of F. nipponica in other Fragaria genomes also possessed more than 70% RNase T2 gene members of their species (72.22% in F. nubicola, 90.48% in F. nilgerrensis, 83.33% in F. viridis, 76.47% in F. vesca, 78.57% in F. iinumae). The chromosomes colinear to chr4 and chr1 in F. nipponica, which are chr2, chr3 and chr6 or chr7 in other genomes, contained no RNase T2 genes in all the six genomes, except for one RNase T2 gene exits on chr7 in F. nubicola genome (Figs. 1B and S4).

Interestingly, the total number of RNase T2 genes in Fragaria nipponica is approximately two-fold of that in F. iinumae. We thus traced the gain-and-loss mechanism for the gene number variation. All species have experienced extensive gene loss, and/or gene number increased via duplication events along with evolution (Fig. 2A). Pseudogenes could also be derived from both non-duplicated genes and duplicated genes (Fig. 2A). F. nipponica showed maximum duplicated genes and minimum loss genes, making its genome contains the highest numbers of RNase T2 genes. In contrast, no duplication event occurred, and a relative high number of gene losses led F. iinumae to retain the fewest RNase T2 genes.

Fig. 2 Copy number variations of RNase T2 homologs. (A) Numbers of identified non-duplicated and duplicated RNase T2 genes in the six Fragaria genomes are shown. (B) Copy numbers of homologous RNase T2 gene sets. Numbers in the rectangles indicated gene copy. HG1-HG12 belongs to class Ⅰ colored by magenta, HG13-HG18 belongs to class Ⅱ colored by dark blue and HG19-HG22 belongs to class Ⅲ colored by green. The phylogeny of species and homologous gene sets (HGs) was inferred from cluster analysis.

We tested whether the biased gain and loss of gene members belonging to any homologous gene set or class (class Ⅰ to Ⅲ) occurred (Fig. 2B). The results showed that HG2, 4, 5, 6, 7 (belonging to class Ⅰ), HG13, 14, 16, 17 (class Ⅱ) and HG21, 22 (class Ⅲ) were lost in at least three Fragaria genomes, which indicated that these HGs tend to be lost in Fragaria, and the gene loss could occur in all three classes. For duplication of these gene family members, HG1, HG3, and HG11 (class Ⅰ) were detected as duplicated genes in at least two Fragaria genomes, among which HG11 was duplicated in four of six Fragaria genomes, but 41.67% of duplicated genes of HG11 were pseudogenes. For HG22, although two copies were identified in the three SI Fragaria genomes, they were actually alleles rather than duplicated genes (see next Result section). Together, the interplay of gene loss, pseudogenization and duplication mainly accounted for the evolution of the RNase T2 gene family size in the Fragaria genomes. To further classify the RNase T2 gene copy disparity among Fragaria species, duplication modes of the multiple-copy genes were investigated. Two types of duplication events were observed, including tandem duplication (TD) and segmental duplication (SD). A chromosomal region within 200 kb containing two or more genes was defined as a tandem duplication event (Holub, 2001). In our study, multiple copies of HG1, HG2 and HG3 were duplicated by TD, and their distances on the chromosome ranged from 7.58 kb (between FvirRNS17_HG3 and FvirRNS18_HG3) to 170.80 kb (between FnubRNS5_HG2 and FnubRNS6_HG2) (Table S3). The two copies of HG5 (class Ⅰ) in the F. nubicola genome and the two copies of HG16 (class Ⅱ) in the F. nipponica genome were duplicated by SD. Additionally, both TD (7 genes, presenting 58.33%) and SD (5 genes, presenting 41.67%) events contributed to copy number variation of HG11 (class Ⅰ) (Table S3). These results imply that TD and SD events play the major driving force on multiple copies of RNase T2 gene members in Fragaria genomes.

3.6. Identification of the candidate T2/S-RNase genes in the Fragaria genomes

Tissue-specific expression is one of the key features of S-RNase genes in SI. To identify the candidate S-RNase genes according to their expression, RNA-seq data of four species (F. nilgerrensis, F. vesca, F. viridis and F. nipponica) from three tissues were obtained, including anther, pistil and mixed tissue (calyx and petal). The results showed that there are four genes belonging to HG22 (FvirRNS1_HG22, FvirRNS2_HG22, FnipRNS9_HG22 and FnipRNS10_HG22) were highly pistil-specific expressed (Fig. 3A and Table S4) and were regarded as candidate S-RNase genes. To further confirm the tissue-specific expression of these four candidate S-RNase genes, RT-PCR was applied in multiple tissues including root, stem, leaf, calyx, petal, anther and pistil. The results showed that the four candidate genes were all specifically expressed in the pistil, which is consistent with the expression features of style determinants (Fig. 3B), providing strong evidence that they are S-RNase genes. An additional candidate gene, FnubRNS14_HG22, was defined based on its homology to the FnipRNS9_HG22 sequence due to unavailable RNA-seq data for F. nubicola. The amino acid sequence similarity of the candidate S-RNase genes ranged from 36.07% to 87.01% (Table S5). Furthermore, their predicted molecular masses are ranged from 26.19 kDa to 29.04 kDa and alkaline isoelectric points (pIs) ranged from 8.49 to 9.29 (Table 1). These features, including phylogenetic relationships (belonging to class Ⅲ), pistil-specific expression, highly polymorphic proteins, a range of molecular masses and alkaline isoelectric points (pIs), are all highly consistent with the known typical S-RNase genes (Igic and Kohn, 2001; Ramanauskas and Igic, 2017). For three self-compatible species, no candidate S-RNase genes were found.

Fig. 3 Transcript expression, RT-PCR experiments and synteny relationship of the S-RNase genes. (A) Expression profiles of the RNase T2 genes in three tissues. The color scale represents relative gene expression levels, and the TPM values are shown inside the rectangle. Gray rectangles represent transcriptome data unavailable for those RNase T2 genes. The five S-RNase genes (FnipRNS9_HG22, FnipRNS10_HG22, FvirRNS1_HG22, FvirRNS2_HG22 and FnubRNS14_HG22) are highlighted with cyan on their branch. (B) RT-PCR experiments verify the gene expression of four S-RNases in multiple tissues including root, stem, leaf, calyx, petal, anther and pistil in Fragaria nipponica and F. viridis. M: Marker (bp). (C) Synteny relationship between two published allelic S-RNases (Sa-RNase and Sb-RNase) and the five newly identified S-RNase genes. Gene locations are highlighted with purple triangles.

Table 1 Genic characteristics of five S-RNase genes identified in the three self-incompatible Fragaria genomes.
Name CDS (bp) Amino acid MW (kDa) pI
FvirRNS1_HG22 762 253 29.04 9.10
FvirRNS2_HG22 753 250 28.50 8.90
FnipRNS9_HG22 744 247 28.73 9.29
FnipRNS10_HG22 696 231 26.78 8.64
FnubRNS14_HG22 696 231 26.49 8.49

Since a recent study identified two allelic S-RNases (Sa-RNase and Sb-RNase) in F. viridis (Du et al., 2021), we aimed to verify whether these two S-RNase genes are the homologous genes to what we identified in this study. Sa-RNase and Sb-RNase sequences were from an assembled genome of another F. viridis accession, and they were both located on chr3 at approximately 468 kb in distance. According to the chromosome collinearity, Sa-RNase and Sb-RNase are orthologous genes with the five candidate S-RNase genes in this study (Fig. 3C). The amino acid sequence similarity between two allelic S-RNases (Sa-RNase and Sb-RNase) and these five S-RNase genes ranged from 34.72% to 92.17% (Table S5). All these results provide solid evidence to indicate that these five candidate genes are S-RNases in three self-incompatible Fragaria species.

3.7. Primary structural features of the candidate T2/S-RNase genes in the Fragaria genomes

The S-RNase genes in Fragaria present exceptional structural variation. Although they all harbored three exons and two introns, the length of the first intron can range from 222 bp to 20.3 kb, and the second intron length ranged from 158 bp to 15.7 kb, which means that both introns showed approximately 100 times variation in length (Fig. 4A). Amino acid sequences of S-RNase genes in Fragaria contain four conserved regions (C1–C3, RC4, and C5) and one hypervariable region (RHV). The first intron is located between signal peptide and C1 while the second intron is within the hypervariable region (Fig. 4B). Structural diversification illustrated that the self-incompatible species evolved fast to accommodate its compatibility.

Fig. 4 Characterization of primary structural features of S-RNase genes. (A) The structure of the S-RNase genes. The genes were drawn to the scale as the scale bar is shown at the bottom, and the lengths of introns are labeled. The published references were highlighted with cyan. Sa/Sb_RNase_Genome: gene structures were drawn according to our unpublished Fragaria viridis genome; Sa/Sb_RNase_published: gene structures were drawn according to the gene sequences published by Du et al., (2021). (B) The five conserved regions (C1–C3, RC4, and C5) and one hypervariable region (RHV) of S-RNase genes.
4. Discussion

One diversifying feature of diploid Fragaria species is that they have evolved both self-compatible and self-incompatible traits (Evans and Jones, 1967; Njuguna et al., 2013; Liston et al., 2014), which have been suggested to be genetically controlled by the S-RNase genes in the RNase T2 gene family (Boskovic et al., 2010; Du et al., 2019, 2021). The phylogenetic analysis in our results also supported that there were multiple transitions from SI to self-compatibility (SC) in Fragaria (Fig. 1). In this study, by using the whole genome sequencing together with the RNA sequencing methods, genome-wide analysis of the RNase T2 gene family in both self-compatible and self-incompatible diploid Fragaria species was performed, and the evolutionary mechanisms of the rapid evolution of T2/S-RNase genes were examined.

Previous analyses of the RNase T2 gene family in other species showed that they were clustered into three classes (Igic and Kohn, 2001; Steinbachs and Holsinger, 2002; Roalson and McCubbin, 2003; Boskovic et al., 2010; MacIntosh et al., 2010; Ramanauskas and Igic, 2017; Du et al., 2019, 2021), but a recent study documented a class Ⅳ RNase T2 gene in seven Rosaceae species (Zhu et al., 2020), among which common introns were absent. In this study, the 115 identified RNase T2 genes were also grouped into three classes (Ⅰ–Ⅲ). Notably, the phylogenetic relationship is largely affected by the correctness of gene annotation. For example, FvesRNS9_HG11 (FvH4_5g24550) and FvesRNS10_HG11 (FvH4_5g24800) in F. vesca involved in class ⅠV were nested into class Ⅰ in our study, indicating that misannotation might lead to classification confusion. Our phylogeny and classification were as robust as in citrus (Honsho et al., 2021) and cacti (Ramanauskas and Igić, 2021).

The identification of RNase T2 gene family members has been performed in diverse plants, and the gene numbers of this family vary among the examined plants. For example, only five RNase T2 genes were found in Arabidopsis (Igic and Kohn, 2001), while there were 19 members in Rubus occidentalis (Zhu et al., 2020). The results need to be closely investigated because the identification was based on the published annotation file, possibly ignoring the un-annotated genes in the genome. For example, there were one to five members that were not found in the annotation file in each examined species in this study (Table S1). Consistent with previous research (Zhu et al., 2020), the quantity in Fragaria nipponica (SI, 27 RNase T2 genes) is approximately twice than that in F. iinumae (SC, 14 RNase T2 genes), but this pattern cannot be applied to all self-incompatible and self-compatible species. This result suggests that total number variations may not be directly related to the transition from SI to SC, but fast-evolving mechanisms existed along with the diversification of Fragaria species.

Gene loss, pseudogenization, and duplication are prevalent in the Fragaria RNase T2 gene family, while duplications, especially TD and SD, are the main evolutionary forces. However, in some other Rosaceae species, RNase T2 duplications are dominated by dispersed duplication (DSD) or whole-genome duplication (WGD) (Qiao et al., 2018), but another case was also reported in pear with TD or PD as the main duplication events (Zhu et al., 2020). Here, we defined only the genes with multiple copies and speculated on the origins of different copies, while other studies evaluated the origin of RNase T2 genes from the entire gene family (Qiao et al., 2018; Zhu et al., 2020). Moreover, the diploid Fragaria species did not experience a recent WGD (Wu et al., 2013; Jiang et al., 2020; Qiao et al., 2021). Therefore, the TD could be a reason for copy variation. Fast evolved gene families can be clearly observed in large gene families but are not very common in small gene family (Meyers et al., 2003; Feng et al., 2017). Therefore, our results provide an example to illustrate the fast evolution in a relatively small gene family.

The Fragaria genus has been proven to engage in the S-RNase-based GSI system (Boskovic et al., 2010; Du et al., 2019, 2021), in which the pistil determination genes are used to prevent self-fertilization (McClure et al., 1989; Sassa et al., 1996; Xue et al., 1996). The distinguishing features of S-RNase genes include specifically high expression in pistils, possessing alkaline pI, and they were classified into class Ⅲ based on the phylogenetic relationship of the entire RNase T2 gene family. The S-RNase genes in potato (Ye et al., 2018), citrus (Liang et al., 2020) and cacti (Ramanauskas and Igić, 2021) were identified by generating style RNA-seq data based on their highly style-specific expression features. In our study, RNA-seq data of styles in four Fragaria species (two self-incompatible and two self-compatible species) were obtained, and ultimately five candidate S-RNase genes (FvirRNS1_HG22, FvirRNS2_HG22, FnipRNS9_HG22, FnipRNS10_HG22 and FnubRNS14_HG22) were found in three self-incompatible species (Fig. 3). These five candidate genes belonging to class Ⅲ are highly pistil-specific (Fig. 3A and B), and their amino acid similarity (36.07%–87.01%) (Table S4), molecular masses (from 26.19 kDa to 29.04 kDa) and alkaline pI (8.49–9.29, Table S1) all fulfill the typical characteristics of the S-RNase gene. Moreover, we confirmed that the two reported allelic S-RNases (Sa-RNase and Sb-RNase) in F. viridis (Du et al., 2021) are the orthologous genes with the five candidate S-RNase genes in this study (Fig. 3C), and high amino acid sequence similarity (92.17%) was observed between Sa-RNase and FvirRNS1_HG22 (Table S4). All these results provide solid evidence to indicate that these five candidate genes are the S-RNases in three self-incompatible Fragaria species, and they may be completely lost after the transition from SI to SC. Surprisingly, the S-RNase genes in Fragaria harbor at least one intron larger than 10 kb (10.4–20.3 kb), except for FnipRNS9_HG22. The intron length also revealed the rapid evolution during the transition from SI to SC. The reason for the exceptionally long intron formation and whether it plays a role in the SI response needs to be further proven.

Acknowledgments

The authors greatly appreciate Dr. Han Guo for enjoyable discussions and providing constructive suggestions to improve this study. We also thank Dr. Jiwei Ruan for his assistance in plant materials collection. This research was financially supported by the National Key Research and Development Program of China (2018YFD1000107) and the open research project of the "Cross-Cooperative Team" of the Germplasm Bank of Wild Species to A.Z.

Authors contributions

A.Z., W.F. and C.Z. designed and conceived this study. W.C., H.W. and A.Z. collected and maintained the plant materials. W.C., H.W., F.L. and H.D. performed the experiments and data analyses. W.C. and W.F. wrote the manuscript with input from A.Z. All authors approved this manuscript.

Declaration of competing interest

The authors declare that they have no conflict of interest.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.pld.2022.04.003.

References
Bailey, T.L., Boden, M., Buske, F.A., et al., 2009. MEME SUITE: tools for motif discovery and searching. Nucleic Acids Res., 37: 202-208. DOI:10.1093/nar/gkp335
Bariola, P.A., Green, P.J., 1997. Plant ribonucleases. In: Ribonucleases: Structure and Functions. Elsevier, pp. 163-190.
Bateman, A., Coin, L., Durbin, R., et al., 2004. The Pfam protein families database. Nucleic Acids Res., 32: 138-141.
Boskovic, R.I., Sargent, D.J., Tobutt, K.R., 2010. Genetic evidence that two independent S-loci control RNase-based self-incompatibility in diploid strawberry. J. Exp. Bot., 61: 755-763. DOI:10.1093/jxb/erp340
Buti, M., Moretto, M., Barghini, E., et al., 2018. The genome sequence and transcriptome of Potentilla micrantha and their comparison to Fragaria vesca (the woodland strawberry). GigaScience, 7: 1-14.
Castresana, J., 2000. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol. Biol. Evol., 17: 540-552. DOI:10.1093/oxfordjournals.molbev.a026334
Chen, C., Chen, H., Zhang, Y., et al., 2020. TBtools: an integrative toolkit developed for interactive analyses of big biological data. Mol. Plant, 13: 1194-1202. DOI:10.1016/j.molp.2020.06.009
Chen, D.L., Delatorre, C.A., Bakker, A., et al., 2000. Conditional identification of phosphate-starvation-response mutants in Arabidopsis thaliana. Planta, 211: 13-22. DOI:10.1007/s004250000271
Deshpande, R.A., Shankar, V., 2002. Ribonucleases from T2 family. Crit. Rev. Microbiol., 28: 79-122. DOI:10.1080/1040-840291046704
Dreesen, R.S.G., Vanholme, B.T.M., Luyten, K., et al., 2010. Analysis of Malus S-RNase gene diversity based on a comparative study of old and modern apple cultivars and European wild apple. Mol. Breed., 26: 693-709. DOI:10.1007/s11032-010-9405-5
Du, J., Ge, C., Li, T., et al., 2021. Molecular characteristics of S-RNase alleles as the determinant of self-incompatibility in the style of Fragaria viridis. Hortic. Res., 8: 185. DOI:10.1038/s41438-021-00623-x
Du, J., Lv, Y., Xiong, J., et al., 2019. Identifying genome-wide sequence variations and candidate genes implicated in self-incompatibility by resequencing Fragaria viridis. Int. J. Mol. Sci., 20: 1039. DOI:10.3390/ijms20051039
Edger, P.P., VanBuren, R., Colle, M., et al., 2018. Single-molecule sequencing and optical mapping yields an improved genome of woodland strawberry (Fragaria vesca) with chromosome-scale contiguity. GigaScience, 7: 1-7.
Emms, D.M., Kelly, S., 2015. OrthoFinder: solving fundamental biases in whole genome comparisons dramatically improves orthogroup inference accuracy. Genome Biol., 16: 1-14. DOI:10.1186/s13059-014-0572-2
Evans, W., Jones, J.K., 1967. Incompatibility in Fragaria. Can. J. Genet. Cytol., 9: 831-836. DOI:10.1139/g67-088
Feng, C., Wang, J., Harris, A.J., et al., 2021. Tracing the diploid ancestry of the cultivated octoploid strawberry. Mol. Biol. Evol., 38: 478-485. DOI:10.1093/molbev/msaa238
Feng, G., Burleigh, J.G., Braun, E.L., et al., 2017. Evolution of the 3R-MYB gene family in plants. Genome Biol. Evol., 9: 1013-1029. DOI:10.1093/gbe/evx056
Finn, R.D., Clements, J., Eddy, S.R., 2011. HMMER web server: interactive sequence similarity searching. Nucleic Acids Res., 39: 29-37.
Folta, K.M., Davis, T.M., 2006. Strawberry genes and genomics. CRC Crit. Rev. Plant Sci., 25: 399-415. DOI:10.1080/07352680600824831
Gasteiger, E., Hoogland, C., Gattiker, A., et al., 2005. Protein Identification and Analysis Tools on the ExPASy Server. In: The Proteomics Protocols Handbook, pp. 571-607.
Grabherr, M.G., Haas, B.J., Yassour, M., et al., 2011. Trinity: reconstructing a full-length transcriptome without a genome from RNA-Seq data. Nat. Biotechnol., 29: 644-652. DOI:10.1038/nbt.1883
Greco, M., Chiappetta, A., Bruno, L., et al., 2012. In Posidonia oceanica cadmium induces changes in DNA methylation and chromatin patterning. J. Exp. Bot., 63: 695-709. DOI:10.1093/jxb/err313
Guindon, S., Dufayard, J.F., Lefort, V., et al., 2010. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst. Biol., 59: 307-321. DOI:10.1093/sysbio/syq010
Hollender, C.A., Geretz, A.C., Slovin, J.P., et al., 2012. Flower and early fruit development in a diploid strawberry, Fragaria vesca. Planta, 235: 1123-1139. DOI:10.1007/s00425-011-1562-1
Holub, E.B., 2001. The arms race is ancient history in Arabidopsis, the wildflower. Nat. Rev. Genet., 2: 516-527. DOI:10.1038/35080508
Honsho, C., Ushijima, K., Anraku, M., et al., 2021. Association of T2/S-RNase with self-incompatibility of Japanese citrus accessions examined by transcriptomic, phylogenetic, and genetic approaches. Front. Plant Sci., 12: 121.
Hu, B., Jin, J., Guo, A.Y., et al., 2015. GSDS 2.0: an upgraded gene feature visualization server. Bioinformatics, 31: 1296-1297. DOI:10.1093/bioinformatics/btu817
Hummer, K.E., Hancock, J., 2009. Strawberry Genomics: Botanical History, Cultivation, Traditional Breeding, and New Technologies. In: Genetics and Genomics of Rosaceae. Springer, New York.
Igic, B., Kohn, J.R., 2001. Evolutionary relationships among self-incompatibility RNases. Proc. Natl. Acad. Sci. U.S.A., 98: 13167-13171. DOI:10.1073/pnas.231386798
Irie, M., 1999. Structure-function relationships of acid ribonucleases: lysosomal, vacuolar, and periplasmic enzymes. Pharmacol. Ther., 81: 77-89. DOI:10.1016/S0163-7258(98)00035-7
Ishimizu, T., Shinkawa, T., Sakiyama, F., et al., 1998. Primary structural features of rosaceous S-RNases associated with gametophytic self-incompatibility. Plant Mol. Biol., 37: 931-941. DOI:10.1023/A:1006078500664
Jiang, S., An, H., Xu, F., et al., 2020. Chromosome-level genome assembly and annotation of the loquat (Eriobotrya japonica) genome. Gigascience, 9: giaa015. DOI:10.1093/gigascience/giaa015
Jiangtao, C., Yingzhen, K., Qian, W., et al., 2015. MapGene2Chrom, a tool to draw gene physical map based on Perl and SVG languages. Hereditas, 37: 91-97.
Jost, W., Bak, H., Glund, K., et al., 1991. Amino acid sequence of an extracellular, phosphate-starvation-induced ribonuclease from cultured tomato (Lycopersicon esculentum) cells. Eur. J. Inorg. Chem., 198: 1-6. DOI:10.1111/j.1432-1033.1991.tb15978.x
Katoh, K., 2002. MAFFT: a novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res., 30: 3059-3066. DOI:10.1093/nar/gkf436
Kim, D., Langmead, B., Salzberg, S.L., 2015. HISAT: a fast spliced aligner with low memory requirements. Nat. Methods, 12: 357-360. DOI:10.1038/nmeth.3317
Köck, M., Theierl, K., Stenzel, I., et al., 1998. Extracellular administration of phosphate-sequestering metabolites induces ribonucleases in cultured tomato cells. Planta, 204: 404-407. DOI:10.1007/s004250050273
Kubo, K.-i., Entani, T., Takara, A., et al., 2010. Collaborative non-self recognition system in S-RNase-based self-incompatibility. Science, 330: 796-799. DOI:10.1126/science.1195243
Kumar, S., Stecher, G., Tamura, K., 2016. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Mol. Biol. Evol., 33: 1870-1874. DOI:10.1093/molbev/msw054
Kurata, N., Kariu, T., Kawano, S., et al., 2002. Molecular cloning of cDNAs encoding ribonuclease-related proteins in Nicotiana glutinosa leaves, as induced in response to wounding or to TMV-infection. Biosci. Biotechnol. Biochem., 66: 391-397. DOI:10.1271/bbb.66.391
Lei, J., Xue, L., Guo, R., et al., 2016. The Fragaria species native to China and their geographical distribution. Acta Hortic., 1156: 37-46. DOI:10.14257/ijsip.2016.9.3.04
Li, Y., Pi, M., Gao, Q., et al., 2019. Updated annotation of the wild strawberry Fragaria vesca V4 genome. Hortic. Res., 6: 61. DOI:10.1038/s41438-019-0142-6
Liang, M., Cao, Z., Zhu, A., et al., 2020. Evolution of self-compatibility by a mutant Sm-RNase in citrus. Nat. Plants, 6: 131-142. DOI:10.1038/s41477-020-0597-3
Liston, A., Cronn, R., Ashman, T.L., 2014. Fragaria: a genus with deep historical roots and ripe for evolutionary and ecological insights. Am. J. Bot., 101: 1686-1699. DOI:10.3732/ajb.1400140
Lu, S., Wang, J., Chitsaz, F., et al., 2020. CDD/SPARCLE: the conserved domain database in 2020. Nucleic Acids Res., 48: 265-268. DOI:10.1093/nar/gkz991
Luhtala, N., Parker, R., 2010. T2 Family ribonucleases: ancient enzymes with diverse roles. Trends Biochem. Sci., 35: 253-259. DOI:10.1016/j.tibs.2010.02.002
MacIntosh, G.C., 2011. RNase T2 family: enzymatic properties, functional diversity, and evolution of ancient ribonucleases. In: Ribonucleases. Springer, pp. 89-114.
MacIntosh, G.C., Hillwig, M.S., Meyer, A., et al., 2010. RNase T2 genes from rice and the evolution of secretory ribonucleases in plants. Mol. Genet. Genomics, 283: 381-396. DOI:10.1007/s00438-010-0524-9
McClure, B.A., Haring, V., Ebert, P.R., et al., 1989. Style self-incompatibility gene products of Nicotlana alata are ribonucleases. Nature, 342: 955-957. DOI:10.1038/342955a0
Meyers, B.C., Kozik, A., Griego, A., et al., 2003. Genome-wide analysis of NBS-LRR-encoding genes in Arabidopsis. Plant Cell, 15: 809-834. DOI:10.1105/tpc.009308
Mota, M., Tavares, L., Oliveira, C.M., 2007. Identification of S-alleles in pear (Pyrus communis L.) cv. 'Rocha' and other European cultivars. Sci. Hortic., 113: 13-19. DOI:10.1016/j.scienta.2007.01.022
Njuguna, W., Liston, A., Cronn, R., et al., 2013. Insights into phylogeny, sex function and age of Fragaria based on whole chloroplast genome sequencing. Mol. Phylogenet. Evol., 66: 17-29. DOI:10.1016/j.ympev.2012.08.026
Ortega, E., Sutherland, B.G., Dicenta, F., et al., 2005. Determination of incompatibility genotypes in almond using first and second intron consensus primers: detection of new S alleles and correction of reported S genotypes. Plant Breed., 124: 188-196. DOI:10.1111/j.1439-0523.2004.01058.x
Pertea, M., Pertea, G.M., Antonescu, C.M., et al., 2015. StringTie enables improved reconstruction of a transcriptome from RNA-seq reads. Nat. Biotechnol., 33: 290-295. DOI:10.1038/nbt.3122
Qiao, Q., Edger, P.P., Xue, L., et al., 2021. Evolutionary history and pan-genome dynamics of strawberry (Fragaria spp.). Proc. Natl. Acad. Sci. U.S.A., 118: e2105431118. DOI:10.1073/pnas.2105431118
Qiao, X., Yin, H., Li, L., et al., 2018. Different modes of gene duplication show divergent evolutionary patterns and contribute differently to the expansion of gene families involved in important fruit traits in pear (Pyrus bretschneideri). Front. Plant Sci., 9: 161. DOI:10.3389/fpls.2018.00161
Ramanauskas, K., Igic, B., 2017. The evolutionary history of plant T2/S-type ribonucleases. PeerJ, 5: e3790. DOI:10.7717/peerj.3790
Ramanauskas, K., Igić, B., 2021. RNase-based self-incompatibility in cacti. New Phytol., 231: 2039-2049. DOI:10.1111/nph.17541
Roalson, E., McCubbin, A.G., 2003. S-RNases and sexual incompatibility: structure, functions, and evolutionary perspectives. Mol. Phylogenet. Evol., 29: 490-506. DOI:10.1016/S1055-7903(03)00195-7
Saint-Oyant, L., Ruttink, T., Hamama, L., et al., 2018. A high-quality genome sequence of Rosa chinensis to elucidate ornamental traits. Nat. Plants, 4: 473-484. DOI:10.1038/s41477-018-0166-1
Sassa, H., Hirano, H., Ikehashi, H., 1993. Identification and characterization of stylar glycoproteins associated with self-incompatibility genes of Japanese pear, Pyrus serotina Rehd. Mol. Gen. Genet., 241: 17-25.
Sassa, H., Nishio, T., Kowyama, Y., et al., 1996. Self-incompatibility (S) alleles of the Rosaceae encode members of a distinct class of the T2/S ribonuclease superfamily. Mol. Gen. Genet., 250: 547-557.
Sonneveld, T., Tobutt, K.R., Robbins, T.P., 2003. Allele-specific PCR detection of sweet cherry self-incompatibility (S) alleles S1 to S16 using consensus and allele-specific primers. Theor. Appl. Genet., 107: 1059-1070. DOI:10.1007/s00122-003-1274-4
Sonneveld, T., Tobutt, K.R., Vaughan, S.P., et al., 2005. Loss of pollen-S function in two self-compatible selections of Prunus avium is associated with deletion/mutation of an S haplotype-specific F-box gene. Plant Cell, 17: 37-51. DOI:10.1105/tpc.104.026963
Steinbachs, J.E., Holsinger, K.E., 2002. S-RNase-mediated gametophytic self-incompatibility is ancestral in eudicots. Mol. Biol. Evol., 19: 825-829. DOI:10.1093/oxfordjournals.molbev.a004139
Taylor, C.B., Bariola, P.A., Delcardayre, S.B., et al., 1993. RNS2: a senescence-associated RNase of Arabidopsis that diverged from the S-RNases before speciation. Proc. Natl. Acad. Sci. U.S.A., 90: 5118-5122. DOI:10.1073/pnas.90.11.5118
Thompson, J.D., Gibson, T.J., Higgins, D.G., 2003. Multiple sequence alignment using ClustalW and ClustalX. Curr. Protoc. Bioinformatics, 1: 2-3. DOI:10.1044/leader.an3.08132003.2
Wang, Y., Tang, H., Debarry, J.D., et al., 2012. MCScanX: a toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Res., 40: e49. DOI:10.1093/nar/gkr1293
Wu, J., Wang, Z., Shi, Z., et al., 2013. The genome of the pear (Pyrus bretschneideri Rehd.). Genome Res., 23: 396-408. DOI:10.1101/gr.144311.112
Xue, Y., Carpenter, R., Dickinson, H.G., et al., 1996. Origin of allelic diversity in antirrhinum S locus RNases. Plant Cell, 8: 805-814.
Ye, M., Peng, Z., Tang, D., et al., 2018. Generation of self-compatible diploid potato by knockout of S-RNase. Nat. Plants, 4: 651-654. DOI:10.1038/s41477-018-0218-6
Zhang, J., Lei, Y., Wang, B., et al., 2020. The high-quality genome of diploid strawberry (Fragaria nilgerrensis) provides new insights into anthocyanin accumulation. Plant Biotechnol. J., 18: 1908-1924. DOI:10.1111/pbi.13351
Zhu, X., Li, Q., Tang, C., et al., 2020. Comprehensive genomic analysis of the RNase T2 gene family in Rosaceae and expression analysis in Pyrus bretschneideri. Plant Syst. Evol., 306: 1-17. DOI:10.1007/s00606-020-01644-0