Skip to main content

ORIGINAL RESEARCH article

Front. Plant Sci., 04 May 2018
Sec. Computational Genomics

Molecular Characterization of SQUAMOSA PROMOTER BINDING PROTEIN-LIKE (SPL) Gene Family in Betula luminifera

  • The State Key Laboratory of Subtropical Silviculture, Institute of Biotechnology, College of Forestry and Biotechnology, Zhejiang A&F University, Lin'an, China

As a major family of plant-specific transcription factors, SQUAMOSA PROMOTER BINDING PROTEIN-LIKE (SPL) genes play vital regulatory roles in plant growth, development and stress responses. In this study, 18 SPL genes were identified and cloned from Betula luminifera. Two zinc finger-like structures and a nuclear location signal (NLS) segments were existed in the SBP domains of all BlSPLs. Phylogenetic analysis showed that these genes were clustered into nine groups (group I-IX). The intron/exon structure and motif composition were highly conserved within the same group. 12 of the 18 BlSPLs were experimentally verified as the targets of miR156, and two cleavage sites were detected in these miR156-targeted BlSPL genes. Many putative cis-elements, associated with light, stresses and phytohormones response, were identified in the promoter regions of BlSPLs, suggesting that BlSPL genes are probably involved in important physiological processes and developmental events. Tissue-specific expression analysis showed that miR156-targeted BlSPLs exhibited a more differential expression pattern, while most miR156-nontargeted BlSPLs tended to be constitutively expressed, suggesting the distinct roles of miR156-targeted and nontargeted BlSPLs in development and growth of B. luminifera. Further expression analysis revealed that miR156-targeted BlSPLs were dramatically up-regulated with age, whereas mature BlmiR156 level was apparently declined with age, indicating that miR156/SPL module plays important roles in vegetative phase change of B. luminifera. Moreover, yeast two-hybrid assay indicated that several miR156-targeted and nontargeted BlSPLs could interact with two DELLA proteins (BlRGA and BlRGL), which suggests that certain BlSPLs take part in the GA regulated processes through protein interaction with DELLA proteins. All these results provide an important basis for further exploring the biological functions of BlSPLs in B. luminifera.

Introduction

SQUAMOSA PROMOTER BINDING PROTEIN-LIKE (SPL) proteins represent a specific transcription factor family in plant. The common characteristic of SPL proteins is the highly conserved DNA-binding domain termed the SBP domain (Yamasaki et al., 2004; Birkenbihl et al., 2005). This domain is comprised of approximately 76 amino acid residues containing two zinc-binding sites, and involved in DNA-binding and nuclear localization (Yamasaki et al., 2004). Due to their ability of binding the promoter region of the floral meristem identity gene SQUAMOSA, SPL genes were firstly identified in Antirrhinum majus (Klein et al., 1996). Lately, SPL genes are identified in all green plant lineages, including green alga, moss, lycophyte, gymnosperm and angiosperm (Riese et al., 2007; Guo et al., 2008; Yang et al., 2008; Li and Lu, 2014; Zhang et al., 2016). And, SPL genes exist as a multigene family in plants. For example, 16 SPL genes are found in Arabidopsis thaliana (Cardon et al., 1999), 19 in rice (Xie et al., 2006; Yang et al., 2008), and 28 in P. trichocarpa (Li and Lu, 2014). Many of SPL genes are important regulators for several diverse plant developmental processes including vegetative phase change (Wang et al., 2009), plant architecture (Jiao et al., 2010; Gou et al., 2017), anthocyanin biosynthesis (Gou et al., 2011), gibberellin (GA) biosynthesis and signaling (Zhang et al., 2007; Yu et al., 2012), somatic embryogenesis (Wang et al., 2017), and in plant response to stresses (Cui et al., 2014).

MiRNAs are a class of noncoding 20–24 nt small RNAs that play important roles in posttranscriptional gene regulation by targeting mRNAs for cleavage or repressing translation in plant (Zhang et al., 2006; Chen, 2009). Many SPL genes are targeted by miR156, which is one of the conserved plant miRNA families. The miR156/SPL module has shown their critical regulatory roles in many development processes (Jiao et al., 2010; Lal et al., 2011; Wang and Wang, 2015). In monocots, the miR156/SPL module has been proposed as a tool to genetically enhance the agronomic traits to maximize the yield in corps (Jiao et al., 2010; Wang and Wang, 2015). For instance, higher expression of OsSPL14, which is regulated by OsmiR156 in vivo, promotes panicle branching in the reproductive stage and increases grain yield in rice (Jiao et al., 2010; Miura et al., 2010). Overexpression of miR156 induced aerial bud formation in switchgrass, while down-regulation of SPL4, one of the miR156 targets, promoted aerial bud formation and increased basal buds, which are closely related to the biomass productivity (Gou et al., 2017). In Arabidopsis, the 3′UTR of AtSPL3 mRNA contains complementary sequences of miR156, and its expression is mediated by miR156 through translation inhibition and transcript cleavage, which defines an important regulatory module of vegetative phase change and floral transition in Arabidopsis (Gandikota et al., 2007; Wang et al., 2009; Yamaguchi et al., 2009). Overexpression of miR156 in transgenic Populus x canadensis decreased the expression of miR156-targeted SPL genes resulting in the severely prolonged juvenile phase, suggesting an evolutionarily conserved role of miR156/SPL module in vegetative phase change of annual herbaceous plants and perennial trees (Wang et al., 2011).

DELLA proteins represent a subgroup of the GRAS family of plant-specific transcription factors, and act as the key negative regulator of GA signaling (Sun, 2010; Locascio et al., 2013). To establish protein-protein interaction with diverse classes of regulatory proteins is the major way to exert function for DELLA proteins (Locascio et al., 2013). By the protein interactions, DELLAs respectively inhibit the DNA-binding ability of transcription factors or the activity of transcriptional regulators, which are important in regulation of various plant development processes (Davière et al., 2014; Rombolá-Caldentey et al., 2014; Huang et al., 2015; Yu et al., 2016). SPL proteins have been revealed as the transcriptional regulators that interact with DELLA proteins. In Arabidopsis, DELLA proteins directly bind to miR156-targeted SPL transcription factors, which disturbs SPL transcriptional activity and consequently delays floral transition (Yu et al., 2012). Therefore, the interaction between SPL and DELLA is thought to be critical for GA pathway in floral transition.

Betula luminifera H. Winkler, a broadleaf tree species, is widely distributed in southern China. Due to its desirable wood properties and fast-growth rate, B. luminifera has been widely grown for timber, being used in high quality furniture, wood veneers and solid wood flooring. In addition to its good economic traits, B. luminifera has a relatively short juvenile period, and many germplasms start flowering within 18 months. Such short vegetative growth phase undoubtedly accelerate the breeding progress of this timber tree, which makes it possible to take B. luminifera as a model for tree breeding.

Although the important roles of SPL genes have been illustrated in model plant, the function of SPL genes in B. luminifera is largely unknown. In the current study, the intron/exon structure, phylogeny, conserved motif, miRNA-mediated posttranscriptional regulation and cis-elements in promoter regions were systematically investigated for SPL genes in B. luminifera. Meanwhile, expression patterns of BlSPLs and mature BlmiR156 in various tissues and organs, and in plants with different ages, were investigated to study the possible roles of BlSPL genes in development, especially in vegetative phase change, of B. luminifera. Moreover, the interactions between BlSPLs and two DELLA proteins were also examined through yeast two-hybrid to study their involvements in the GA regulated biological processes. Our findings provide a solid foundation for future research to elucidate the functions of the SPL genes in B. luminifera, and will surely promote the utilization of these genes.

Materials and Methods

Plant Materials

Plant samples from Betula luminifera (clone 1V25-2), which grown in the Tree Germplasm Garden of Zhejiang A&F University, were used for gene cloning and expression analysis. For the gene cloning, stems, leaves, and inflorescences were collected from 3-year-old plants in April, and then stored in liquid nitrogen until use. For expression analysis in different tissues and organs, buds, leaves, male inflorescences, female inflorescences, stems, root, phloem, xylem, and seeds were collected from 3-year-old plants between early March and Late May. For expression analysis in seedling/plants with different ages, the seedlings germinated from self-pollinated seeds of clone 1V25-2, and the 3-, 5- month-old, 1.5- and 4-year-old trees of clone 1V25-2 were used. The 5-, 10-, and 21-day-old seedlings being excised roots, and the 2-3 cm newly-expanded leaves from the shoot apex of trees were collected and stored in liquid nitrogen until use.

Identification of BLSPL Genes

The amino acid sequences of AtSPLs were retrieved from the Arabidopsis Information Resource database (TAIR version 10, http://www.arabidopsis.org/), and the sequences of Populus PtSPLs were obtained from the previous study (Li and Lu, 2014). To identify the genomic sequences (contigs) harboring BlSPL genes, the AtSPL and PtSPL protein sequences were used for TBLASTN search (Altschul et al., 1997) against the current assembly of the B. luminifera genome (Huang et al., unpublished data 2017) with the parameter “-evalue 1e-5”. Then, the candidate contigs identified by TBLASTN (Supplementary Table S1) were used to predict the gene models of BlSPLs using the FGENESH online tool with the genome-specific gene-finding parameters of A. thaliana and P. trichocarpa (Softberry, http://www.softberry.com/berry.phtml; Solovyev et al., 2006). All the nonredundant gene models were submitted to Interpro (http://www.ebi.ac.uk/interpro) to confirm the presence of the SBP domain. Based on the predicted models, all the BlSPL genes were cloned and sequenced, and the predicted gene models were then further manually corrected and verified by alignment with the cloned sequences. The predicted and corrected gene models were listed in the Supplementary Table S2. All the sequences of BlSPL genes were deposited in Genbank (https://www.ncbi.nlm.nih.gov/genbank/) and the accession numbers were listed in Table 1.

TABLE 1
www.frontiersin.org

Table 1. Gene feature and classification of SPLs in Betula luminifera.

Phylogenetic Analysis

Multiple sequences alignment of SPL proteins was performed using AlignX to obtain SBP domain sequences. Phylogenetic analysis was carried out by the neighbor-joining (NJ) method in MEGA 7.0 with 1,000 bootstrap replicates (Kumar et al., 2016).

Bioinformatic Analyses of BLSPL Genes

The theoretical isoelectric point (pI) and molecular weight (Mw) of BlSPLs were calculated using the Compute pI/Mw tool on the ExPASy server (http://web.expasy.org/compute_pi/). Intron/exon structures were displayed by mapping the cDNA sequences of BlSPLs to the corresponding genomic sequences using the Gene Structure Display Sever 2 (Hu et al., 2015; http://gsds.cbi.pku.edu.cn/). Conserved domains of BlSPLs were identified using Pfam (http://pfam.sanger.ac.uk) and the Conserved Domain Database (CDD, http://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi) with the expected E-value threshold of 0.01 and the maximum hits of 500 amino acids (Zhang et al., 2013). Sequence logos were generated by using WebLogo (http://weblogo.berkeley.edu/). Sequence alignment of SBP domains, ANK domains and miR156 complementary sequences was carried out using Vector NTI 11.0 (Invitrogen). Conserved motifs were determined using MEME version 4.11.4 (Bailey and Elkan, 1994) (http://meme-suite.org/tools/meme) with the parameters as previous study (Li and Lu, 2014). MiR156-targeted BlSPLs were predicted using psRNATarget tool (http://plantgrn.noble.org/psRNATarget/; Dai and Zhao, 2011).

Cis-Elements Analysis

Putative cis-elements in the promoter regions (1,500 bp upstream sequences of the start codon) of BlSPLs were annotated using PlantCARE (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/) database (Lescot et al., 2002). The motifs putatively involved in plant growth and development, phytohormone responses, and stress responses are summarized.

RNA Isolation

Total RNA was extracted from tissues or organs of B. luminifera using the PureLink Plant RNA Reagent (Ambion) according to the manufactuer's instructions. Agarose gel electrophoresis and nanodrop 2000 spectrophotometer (Thermo Scientific) were used to evaluate the quality and quantity of total RNA, and genomic DNA was eliminated by treatment of RNase-free DNase (Promega).

Molecular Cloning of BLSPLs

According to SMARTer RACE cDNA Amplification Kit (ClonTech), the full length of BlSPLs were obtained by RACE method using cDNA of Betula luminifera as template by PCR, and using the gene specific forward and reverse primers listed in Supplementary Table S3. The complete coding regions of BlSPLs were amplified and verified by PCR using the specific primers listed in Supplementary Table S3. PCR products were purified, cloned into pMD-19 T vector (Takara), and then were sent to sequencing (Genscript, Nanjing).

Experimental Validation of Cleavage of miR156-Targeted BLSPLs

To identify the cleavage sites of miR156-targeted BlSPLs, the modified RNA ligase-mediated (RLM) rapid amplification of 5' cDNAs method (5' RLM-RACE) was performed using the First Choice RLM-RACE Kit (Invitrogen) as described previously (Guo et al., 2005). Nested gene specific primers were designed according to the predicted cleavage sites and are listed in Supplementary Table S4. The 5'-RACE products were purified and cloned into pEAST-T1 vector (TransGen Biotech, Beijing). At least eight clones were taken and confirmed by Sanger sequencing.

Expression Analysis of BLSPL and DELLA Genes

The cDNA was synthesized according to the manufacturer's instructions of PrimeScript RT reagent Kit (Takara). The quantitative reverse transcription-PCR (qRT-PCR) was performed on the CFX96 thermocycler (Bio-Rad) using SYBR Green qPCR Master Mix (Takara). The program used for qRT-PCR is as follows: pre-denaturation at 95°C for 30 s, 40 cycles of amplification at 95°C for 5 s, annealing at 60°C for 20 s and 72°C for 15 s. Elongation factor 1-alpha gene (EF1α) (Genbank accession no. KM586061) was used as a reference gene as described previously (Liu et al., 2016). Melting curve was used to evaluate amplification specificity. The relative expression levels of BlSPL and DELLA genes were analyzed using the 2−ΔΔCq method (Schmittgen and Livak, 2008). Normalization of gene expression data from three biological replicates was performed as described (Willems et al., 2008). All the primers used for qRT-PCR are listed in Supplementary Table S5.

Expression Analysis of BlmiR156

Expression of BlmiR156 was analyzed using specific reverse transcription primers as the method described in the previous studies (Varkonyi-Gasic et al., 2007; Li et al., 2013). Briefly, 200 ng total RNA treated by DNase was used for first-strand cDNA synthesis using PrimeScript 1st Strand cDNA Synthesis Kit (TaKaRa). A reaction volume of 10 μL was set up according to the manufacturer's instructions, using 0.5 μL miR156 reverse transcription primer (10 mM) and 0.1 μL 5.8S rRNA reverse primer (10 mM) instead of Poly dT and random hexamer primers. The qRT-PCR was performed as described above. The 5.8S rRNA (GenBank accession no. KT308944) was used as a reference gene in the qRT-PCR. The primers used in the expression analysis of BlmiR156 are listed in Supplementary Table S5.

Yeast Two Hybrid Assay

The yeast strains and media used for the yeast two-hybrid experiment were provided in the Matchmaker Golden Yeast Two-Hybrid System (Clontech). The open reading frames (ORF) of BlSPL genes, BlRGA (Genbank accession no. MF149049) and BlRGL (Genbank accession no. MF149050) were amplified and firstly cloned into pCR™8/GW/TOPO vector (Invitrogen). Then, these ORFs were cloned into pDEST-GBKT7 (GAL4 DNA binding domain, BD) or pDEST-GADT7 (GAL4 activation domain, AD; Rossignol et al., 2007) through LR recombination reaction (Invitrogen). The BD fused constructs and AD fused constructs were transformed into Y2HGold cells and Y187 cells by the lithium acetate-mediated method, respectively. The yeast two-hybrid assay was performed by yeast mating according to the user manual. The BD fused constructs and AD fused constructs were tested for their autoactivation and toxicity, and those constructs without autoactivation and toxicity in yeast cells were used for yeast two-hybrid assay (Supplementary Figure S1). To measure the transcription activation activity, β-Galactosidase activity was assayed using o-nitrophenyl β-D-galactopyranoside (ONPG) as described in Yeast Protocol Handbook (Clontech). The primers used for ORF amplification are listed in Supplementary Table S6.

Results

Identification, Molecular Cloning and Gene Feature Analysis of BLSPLs

In order to identify SPL genes in B. luminifera, all SPL proteins of Arabidopsis and Populus were used to query against the draft genome (Huang et al, unpublished data, 2017) using TBLASTN algorithm (Altschul et al., 1997). The genomic sequences identified by TBLASTN were then used to predict gene models by using the FGENESH online tool. As shown in the Supplementary Table S2, 18 gene models containing complete SBP-domain were obtained. These predicted gene models were further verified and corrected through molecular cloning and sequencing. The corrected gene models indicated the existence of 18 BlSPL genes in genome of B. luminifera, and were named BlSPL1 to BlSPL18 (Table 1, Supplementary Table S2).

Further analysis of the experimentally validated cDNA sequence of BlSPLs showed that the lengths of BlSPL cDNAs varied between 720 and 4,234 bp, and the deduced protein lengths from 134 to 1,082 amino acids (Table 1). The theoretical pI of deduced BlSPL proteins ranged from 5.89 to 9.34, and the molecular weight (Mw) varied from 15.4 to 119.4 kDa (Table 1). In addition, the number of introns varied between 1 and 9 in BlSPL genes (Table 1, Figure 3), which is similar with the number of AtSPL and PtSPL genes. These results revealed the diversity of SPL genes features in B. luminifera.

Analyses of Conserved Domain, Phylogeny, and Intron/Exon Structure

The detailed domain structures of BlSPLs were analyzed by multiple sequence alignment. As the result, SBP domain was the only one conserved domain found to be shared by all BlSPLs (Figure 1A). High similarity of the SBP domains, especially several extremely conserved positions, was observed in all BlSPLs (Figure 1B). All of the SBP domains in BlSPLs shared two zinc finger-like structures (Zn-1, Zn-2). The Zn-1 motif was C3H (CCCH) type in all the SPL proteins except BlSPL10, in which the His residue was replaced by Cys residue (Figure 1A). Unlike Zn-1, the signature sequence (C2HC) of Zn-2 is highly conserved in all BlSPLs. Except for the zinc finger-like structure, all the BlSPLs contain a conserved nuclear location signal (NLS) in the C-terminus of SBP domains, which was partly overlapped with the Zn-2 motif (Figure 1A). These structure features of SBP domains were also observed in other plant SPLs (Yang et al., 2008; Li et al., 2013; Li and Lu, 2014; Zhang et al., 2016), indicating the highly conservation of SBP domain in plants. Moreover, three BlSPLs (BlSPL4, BlSPL5, and BlSPL12) contain an ANK or Ank-2 domain with four or three ankyrin repeats (Supplementary Figure S2), which were less conserved than SBP domain and believed to be associated with protein-protein interaction (Michaely and Bennett, 1992).

FIGURE 1
www.frontiersin.org

Figure 1. SBP-domain alignment of BlSPLs. (A) Multiple alignment of the SBP domains of the B. luminifera SBP-box proteins was obtained with the Vector NTI Advanced 11 software. The two conserved zinc finger structures (Zn-1, Zn-2) and NLS were indicated. (B) Sequence logo of the SBP domain of BlSPLs. The overall height of each stack represents the degree of conservation at this position, while the height of the letters within each stack indicates the relative frequency of the corresponding amino acids.

In order to investigate the evolutionary relationship of BlSPLs with SPL genes from other plants, 81 SPL sequences were collected to construct an unrooted phylogenetic tree using neighbor-joining (NJ) method in MEGA7.0. These 81 SPLs were selected from four plant species, including 18 from B. luminifera, 16 from Arabidopsis thaliana, 19 from Oryza sativa (rice), and 28 from P. trichocarpa (Supplementary Tables S7, S8). Only the highly conserved SBP domains (76 aa) sequences were used for phylogenetic analysis (Riese et al., 2007; Li et al., 2013; Zhang et al., 2016). As a result, these SPL proteins were clustered into 9 groups (I-IX), each of which included at least one BlSPL (Figure 2). Remarkably, all the BlSPL genes, except BlSPL14, that clustered into group IV to IX were putative target genes of BlmiR156 (Table 1 and Figure 2). As reported in previous studies, this phylogenetic tree indicated the evolutionary diversification and differentiation of BlSPL genes as well as other plant SPL genes.

FIGURE 2
www.frontiersin.org

Figure 2. Phylogenetic analysis of B. luminifera and other plant SPLs. Phylogenetic tree was constructed with SBP domain protein sequences from B. luminifera (BlSPL), rice (OsSPL), P. trichocarpa (PtSPL), and Arabidopsis (AtSPL). SPL genes from B. luminifera that contain complementary sequences to mature BlmiR156 are marked with asterisks. SBP domain sequences of all genes used for phylogenetic tree construction, as well as the accession numbers or locus IDs and data sources of SPL genes from different species are listed in Supplementary Tables S7, S8.

By aligning the coding sequences to corresponding genomic sequences, intron/exon structures of all 18 BlSPL genes were also analyzed (Figure 3). As shown in Figure 3, high variation in the number of introns was revealed, which ranges from one (BlSPL3, BlSPL14, BlSPL15, and BlSPL17) to 9 (BlSPL4, BlSPL5, BlSPL10, and BlSPL12). Based on the BlSPL tree (Figure 3A), group I and II proteins contain 9 introns, group III and IV contain 2, group V contains 3, group VI contains 1, group VII, and VIII contain 2, group IX contains 3 introns (Figure 3B). In general, the BlSPL genes shared a similar intron/exon structure within the same group. For example, BlSPL genes within group VI had only one intron, whereas those within group II contained nine introns. In addition, all the SBP domains of BlSPL genes were separated by the first intron, which were highly variable in length among the BlSPL genes.

FIGURE 3
www.frontiersin.org

Figure 3. Phylogenetic analysis and intron/exon structures of BlSPL genes. (A) A phylogenetic tree was constructed with B. luminifera SBP domain protein sequences. (B) Exons and introns are indicated by green boxes and black horizontal lines, respectively.

Identification of Conserved Motifs in BLSPLs

By using the MEME web server, a total of 12 conserved motifs for 18 BlSPLs were identified (Figure 4, Table 2). The number of motifs in each BlSPL varies from 1 to 8, and most BlSPLs shared similar motif profiles within the same group (Figure 4). Among these motifs, motif 1 is actually the SBP domain, which exists in all BlSPLs analyzed. Motif 6 existed in Group VI-V and VII-IX BlSPLs contains the complementary sequence of miR156, suggesting the posttranscriptional regulation of Group VI-V, VII-IX BlSPLs by miR156. In addition, motif 8 also widely exists in several SPL groups. Interestingly, several motifs are found to be group-unique, such as motif 12 specifically existing in group VII, and motifs 3 (ANK domain) and 7 are found exclusively in group II (Figure 4). Functions of these group-unique motifs are currently unclear, but should be tightly related to specific roles of SPLs in the group.

FIGURE 4
www.frontiersin.org

Figure 4. Distribution of conserved motifs in BlSPLs. Motifs represented with boxes are identified using MEME online tool. The number in boxes (1–12) represents motif 1-motif 12, respectively. The position and length of each colored box represents the actual motif size.

TABLE 2
www.frontiersin.org

Table 2. Consensus sequences of 12 motifs identified in BlSPLs.

Posttranscriptional Regulation of BLSPLs Mediated by miR156

It has been confirmed that a subset of SPL genes are regulated by miR156 at posttranscriptional level in plant (Schwarz et al., 2008; Wang et al., 2009, 2011; Jiao et al., 2010; Wang and Wang, 2015; Tripathi et al., 2017). With the psRNATarget online tool, a total of 12 BlSPLs were predicted to be targeted by miR156 (Figure 5).The complementary sites of miR156 are in coding regions for nine BlSPLs (BlSPL1, BlSPL6, BlSPL7, BlSPL8, BlSPL9, BlSPL11, BlSPL13, BlSPL16, and BlSPL18) belonging to group IV, V, VII, VIII, and IX, whereas it locates in the 3'-UTR of BlSPL3, BlSPL15 and BlSPL17 within group VI (Figure 5D). Sequence alignment indicated that the miR156 complementary sites were quite conserved across these BlSPL genes (Figure 5A), and the sequence divergence was mainly restricted to the first, second and last nucleotide of complementary sequences (Figure 5B). It means that the complementary sites are actually under strong selection pressure during the evolution, even for those locating in the 3'-UTR. These results are consistent with the previous studies from other plants, suggesting the importance of miR156-mediated posttranscriptional regulation for functions of SPL genes in plant.

FIGURE 5
www.frontiersin.org

Figure 5. Analysis of miR156-mediated posttranscriptional regulation of BlSPLs. (A) Sequence alignment of miR156 complementary sequences of the BlSPLs. The 50 nucleotides upstream and downstream of the miR156 complementary sequences were included in the alignment. (B) Sequence logo of the miR156 complementary sequences of BlSPL genes. (C) Confirmation of the cleavage sites in twelve BlSPL transcripts by RLM 5'-RACE. Vertical arrows indicate the cleaved nucleotide position, with the frequency of clones shown. (D) Distances between important sites in sequences of twelve BlSPL genes. Important sites include start codon, SBP domain, and binding sites, cleavage sites of the miR156 complementary sequences with their nucleotide positions. Heavy gray lines represent open reading frames (ORFs). The lines flanking ORFs represent 3'-UTR. SBP domains are indicated in blue. MiR156 complementary sequences of BlSPL genes are indicated in green. Distances between these sites are measured and indicated by base pair (bp).

In order to validate the cleavage of the BlSPLs transcript by miR156, RLM 5'-rapid amplification of the cDNA ends (5'-RACE) was performed for all twelve BlSPLs. The result indicated that two cleavage sites of miR156 were detected in the complementary sequences of these twelve BlSPLs (Figure 5C). One of the cleavage sites was present between the 10th and 11th nucleotides of the complementary sequences in ten BlSPLs (BlSPL1, BlSPL3, BlSPL6, BlSPL7, BlSPL9, BlSPL11, BlSPL13, BlSPL15, BlSPL17, and BlSPL18), while the other cleavage site was located between 11th and 12th nucleotides of the complementary sequences in three BlSPLs (BlSPL1, BlSPL8 and BlSPL16) (Figure 5C). Interestingly, both of these two cleavage sites were detected in BlSPL1. Thus, all these results validate the posttranscriptional regulation of the BlSPL genes by miR156 in B. luminifera, and imply that these BlSPLs might be regulated by different miR156 genes through different cleavage sites.

Moreover, by comparing the distances between important sequence sites of these BlSPL genes, three types of distance between SBP domain and complementary site were revealed. It could be roughly classified into short-distance (less than 200 bp), medium-distance (approximately 350 bp) and long-distance (more than 400 bp) (Figure 5D). Similar distance type was observed for the BlSPL genes within the same phylogenetic groups. For instance, the BlSPL3, BlSPL15, and BlSPL17 in group VI have the short-distance type, whereas the BlSPLs of group VII have the long-distance type (Figure 5D). This implies that a common ancestral gene may have experienced different evolutionary events resulting in the differentiation of these miR156-regulated BlSPL genes.

Cis-Elements in the Promoter Regions of BLSPL Genes

Cis-elements play important roles in the regulation of gene transcription during plant growth, development, and stress responses. To understand the transcriptional regulation mechanisms, the cis-elements in promoter regions of BlSPL genes were identified through PlantCARE database. Except for the common cis-acting elements such as CAAT-box and TATA-box, many cis-elements were identified in promoter regions of 18 BlSPL genes (Supplementary Table S9). According to their putative functions, these elements were categorized into eight classes, which are shown in the Figure 6. As the results, light-responsive elements had the largest number and were present in all BlSPL gene promoter regions (Supplementary Table S9, Figure 6). And, the hormone responsive elements, plant tissue-specific elements, and stress responsive elements are also found to be present in promoter regions of all BlSPL genes (Supplementary Table S9, Figure 6). In addition, other rarely distributed cis-elements in BlSPLs were found to be functionally involved in transcription regulation, circadian control, protein binding, elicitor responsiveness (Supplementary Table S9, Figure 6). Thus, the transcription of BlSPL genes could be regulated by various environmental and developmental changes, which implied that BlSPL genes were involved in important physiological processes and developmental events. Furthermore, there no very similar cis-elements distribution was observed among these BlSPL genes, even for those BlSPLs in the same phylogenetic group (Supplementary Table S9, Figure 6). These observations suggest that the differentiations in the promoter regions may promote the neofunctionalization of BlSPL genes during their divergence and evolvement.

FIGURE 6
www.frontiersin.org

Figure 6. Cis-elements in promoter regions of 18 BlSPLs. Elements were identified in the 1500 bp sequences upstream of the start codon of BlSPL genes using PlantCARE database. The elements associated with specific functions are indicated by different colors for each gene.

Expression Profiles of the BLSPL Genes and Blmir156 in Different Tissues/Organs of B. luminifera

In order to explore the possible roles of BlSPLs in development of B. luminifera, the expression profiles of these 18 BlSPL genes in various tissues/organs were investigated using the quantitative real-time RT-PCR method (qRT-PCR). In general, the expression patterns of BlSPL genes can be classified into two types according to their expression profiles (Figure 7). The minority of BlSPL genes, including BlSPL4, BlSPL5, BlSPL10, and BlSPL12, were constitutively expressed in all tissues/organs examined (Figure 7B). And, all these genes are BlSPL genes without miR156 complementary sites (termed as BlmiR156-nontargeted BlSPL genes). The remaining majority of BlSPL genes, most of which are BlSPL genes with miR156 complementary sites (termed as BlmiR156-targeted BlSPL genes), exhibited a more differentiated expression pattern in different tissues or organs (Figure 7A). This difference implies distinct roles of BlmiR156-targeted and nontargeted BlSPL genes in development of B. luminifera. In addition, similar expression patterns were observed among those genes belonging to same group, such as BlSPL6/7 in group V, BlSPL4/5/12 in group II, BlSPL14/15 in group VI (Figure 7), suggesting the redundant functions of these BlSPL genes. On the other hand, it was particularly worthy to note that many BlSPLs were expressed at a relatively low level in seeds compared with their expressions in other tissues or organs analyzed (Figures 7A,B). Furthermore, the transcript levels of mature BlmiR156 in different tissues or organs of B. luminifera were also investigated by qRT-PCR. According to its expression pattern, the mature BlmiR156 was mainly accumulated in seeds, which was approximately four times of the second highest level in male inflorescences at late stage (MI3) (Figure 7C). In contrast, in many tissues or organs, such as the buds, young leaves (L1 and L2), female inflorescences, stems, phloem and xylem, the transcript levels of mature BlmiR156 were rather low (Figure 7C).

FIGURE 7
www.frontiersin.org

Figure 7. Expression profiles of BlSPLs and mature BlmiR156 in different tissues of B. luminifera. (A) Expression patterns of BlmiR156-targeted BlSPLs. (B) Expression patterns of BlmiR156-nontargeted BlSPLs. (C) Expression patterns of BlmiR156. Fold changes of transcript levels in buds (Bd), unexpanded leaves (L1), newly expanded leaves (L2), mature leaves (L3), male inflorescence at early stage (MI1), male inflorescence at mid stage (MI2), male inflorescence at late stage (MI3), female inflorescence at early stage (FI1), female inflorescence at mid stage (FI2), female inflorescence at late stage (FI3), young stems (St1), mature stems (St2), root (R), phloem (Ph), xylem (Xy), Seed (Sd) of Betula plants are shown. Transcript levels in buds were arbitrarily set to 1 and the levels in other tissues were given relative to this. Error bars represent standard deviations of mean value from three biological replicates and four technical replicates.

Expression Profiles of the BLSPL Genes and BlmiR156 With Age in B. luminifera

MiR156 regulated SPL genes play vital roles in regulation of plant vegetative phase change (Wang et al., 2009, 2011). To investigate the roles of BlSPL genes in vegetative phase change of B. luminifera, the expression patterns of BlSPL genes and mature BlmiR156 were examined in seedlings or fully expanded leaves of plants with different ages. Similar as in different tissues or organs, two distinct expression patterns of these BlSPL genes were observed (Figure 8). Most of the BlmiR156-targeted BlSPL genes, especially for BlSPL1, BlSPL3, and BlSPL6, were dramatically up-regulated with age in plants (Figure 8A), while most of the BlmiR156-nontargeted BlSPL genes showed a constitutive expression pattern (Figure 8B). In contrast, the mature BlmiR156 was strongly accumulated in 5- and 10-day-old seedlings, and dramatically down-regulated in 1.5-year-old and 4-year-old plants (Figure 8C). These results indicate the functional differentiation between BlmiR156-targeted and nontargeted BlSPLs, and also suggest important roles of these BlmiR156-targeted BlSPL genes in vegetative phase change of B. luminifera.

FIGURE 8
www.frontiersin.org

Figure 8. Expression profiles of BlSPLs and mature BlmiR156 with age in B. luminifera. (A) Expression patterns of BlmiR156-targeted BlSPLs. (B) Expression patterns of BlmiR156-nontargeted BlSPLs. (C) Expression patterns of BlmiR156. Fold changes of transcript levels in 5- (5 d), 10- (10 d) and 21-day-old (21 d) seedlings, and 3- (3 m), 5- (5 m), 1.5-year (1.5 y), and 4-year-old (4 y) plants are shown. Transcript levels of BlSPLs in 5-day-old (5 d) seedlings were arbitrarily set to 1 and the levels in other tissues were given relative to this. Transcript levels of BlmiR156 in 4-year-old (4 y) plants were arbitrarily set to 1 and the levels in other samples were given relative to this. Error bars represent standard deviations of mean value from three biological replicates and four technical replicates.

Protein Interaction Between BLSPLs and DELLA Proteins

The interactions between miR156-targeted SPL transcription factors and DELLA proteins have been suggested to be an important regulatory module for GA regulated floral transition in plant (Yu et al., 2012; Yamaguchi et al., 2014). In our study, yeast two-hybrid assay was used to survey the interaction between BlSPL proteins and two DELLA proteins (BlRGA and BlRGL) of B. luminifera. As shown in Figure 9, similar interaction patterns were observed between BlSPLs and these two DELLA proteins. There six BlSPL proteins (BlSPL1/2/5/8/13/18) interact with both DELLA proteins, while the BlSPL6 only interact with BlRGA (Figure 9A), and BlSPL16 only interact with BlRGL (Figure 9B). It is worth to note that not only BlmiR156-targeted BlSPL genes but also nontargeted BlSPL genes (BlSPL2 and BlSPL5) could interact with DELLA proteins. On the other hand, the interaction strength between these BlSPL proteins and DELLA proteins are obvious different. For example, BlSPL1, BlSPL8, and BlSPL18 tightly interact with BlRGA and BlRGL. Conversely, the interaction between BlSPL2 and BlRGA as well as BlSPL16 and BlRGL are very weak (Figure 9). The strength differences between these interactions were also validated through β-galactosidase activity assay (Supplementary Figure S3). These results indicate interaction with DELLA proteins is possibly significant for functions of these BlSPL genes. And, the BlmiR156-targeted BlSPL genes may also participate in GA regulated flowering in B. luminifera.

FIGURE 9
www.frontiersin.org

Figure 9. Interactions of BlSPLs with DELLA proteins. (A) Interaction between BlSPLs and BlRGA revealed by yeast two-hybrid. (B) Interaction between BlSPLs and BlRGL revealed by yeast two-hybrid. BD, GAL4 binding domain; AD, GAL4 activation domain. Interactions were examined on the SD-Leu-Trp-His-Ade plates supplemented with 40 μg/ml X-α-gal and 125 ng/ml Aureobasidin A (SD-L-W-H-A-X-AbA). Yeast grown on the SD medium without leucine and tryptophan (SD-L-W) were used as a control. The yeast clones were grown with dilutions to 10−1 and 10−2.

Additionally, the expressions of these two DELLA genes were also investigated in different tissues/organs and plants with different ages. As shown in Supplementary Figure S4, different expression patterns were revealed for these two DELLA genes. The BlRGA was highly expressed in xylem (Xy) and male inflorescence at early stage (MI1) (Supplementary Figure S4A), and reached to highest expression level in 21-day- old plants (Supplementary Figure S4C). Whereas, the BlRGL was evenly expressed in most tissues/organs examined (Supplementary Figure S4B), and gradually up-regulated with age in plants (Supplementary Figure S4D). For those interacted BlSPL genes, further correlation analysis indicated that the expression of BlSPL5 was significantly correlated with the expressions of BlRGA and BlRGL in different tissues/organs, while the expressions of BlSPL1, BlSPL8, and BlSPL13 were merely significantly correlated with that of BlRGL in plants with different ages (Supplementary Table S10). It means that these BlSPL genes might be co-expressed with the DELLA genes in specific tissue/organ or developmental stage, which may determine their protein interactions and functions.

Discussion

SPLs are one family of the plant-specific transcription factors featured with a highly conserved SBP domain, which is capable of binding to the promoters of floral meristem identity gene SQUAMOSA and its homologs (Klein et al., 1996). The SPL genes have been identified from many plant species, such as Arabidopsis (Cardon et al., 1999), rice (Yang et al., 2008), soybean (Tripathi et al., 2017), Populus (Li and Lu, 2014), chrysanthemum (Song et al., 2016), and pepper (Zhang et al., 2016). However, no genome-wide identification of SPL gene family has been reported in B. luminifera. In this study, genome-wide identification of the SPL gene family was performed in B. luminifera for the first time. As a result, 18 SPL genes were identified in the genome of B. luminifera, which is similar to the number of SPL genes in Arabidopsis (16) and rice (19), but much less than that of P. trichocarpa (28), or that of moso bamboo (32) (Pan et al., 2017). It implies that the SPLs of different plants evolved in a species-specific manner, which is possibly influenced by different gene duplication events.

Based on the phylogenetic analysis, the BlSPL gene family was divided into nine groups (I-IX) (Figure 2). And, the BlSPL genes grouped tightly with AtSPL and PtSPL genes, reflecting the fact that these SPL genes may diverge more recently from a common ancestor. Gene structure analyses showed that BlSPL genes shared similar intron/exon structures within the same phylogenetic group (Figure 3), as described previously in cotton (Cai et al., 2018), apple (Li et al., 2013), tomato (Salinas et al., 2011), and rice (Xie et al., 2006). Moreover, most BlSPLs of the same phylogenetic group also shared similar motifs (Figure 4), which showed that these genes within the same phylogenetic group may have similar functions in B. luminifera. Interestingly, except for the conserved motifs present in most BlSPLs, several unique motifs were also found in specific phylogenetic groups (Figure 4), such as motif 4 in group II, motif 10 in group V and motif 12 in group VII. These unique motifs are probably important for their specified functions, indicating that the many BlSPL genes may have undergone functional differentiation and neofunctionalization with the divergence of different lineages.

MiR156-mediated posttranscriptional regulation has been considered as an important mechanism for the functions of a subset of SPLs (Gandikota et al., 2007; Wang et al., 2009, 2017; Gou et al., 2011, 2017; Kim et al., 2012). In Arabidopsis, 10 of 17 AtSPLs are targets of miR156 (Rhoades et al., 2002). Also, 11 of 19 SPLs in rice (Xie et al., 2006), 18 of 28 SPLs in Populus (Li and Lu, 2014), 10 of 15 SPLs in Citrus (Shalom et al., 2015) and 12 of 18 SPLs in Ziziphus jujube (Shao et al., 2017), were reported to be targeted by miR156. In our study, target prediction and experimental validation showed that 12 BlSPLs in groups IV-IX were regulated by miR156 (Figure 5). In addition, two kinds of the complementary sites of miR156 were detected in these BlSPLs, one of which locates in 3' UTR of BlSPLs in group VI, whereas the other one locates in the coding region. High conservation of the miR156 complementary sequences across all miR156-targeted BlSPLs was also observed in both coding regions and 3' UTR (Figures 5A,B). It is consistent with the results from SPLs of Arabidopsis, Salvia miltiorrhiza, Populus, Ziziphus jujuba and Citrus, which suggests that the miR156-mediated posttranscriptional regulation of SPL genes should be highly conserved in plants.

Cis-elements in the promoter regions are closely correlated with gene expression and/or tissue specificity in development and responses to stress. Many cis-elements were found in the promoter regions of BlSPLs (Figure 6, Supplementary Table S9), and lots of these cis-elements are associated with light-responsiveness, hormone responsiveness, transcription regulation and stresses responses. Similar cis-elements were also identified in moso bamboo and upland cotton, which suggests that the expressions of SPL genes may be regulated by light, stresses and/or phytohormones (Pan et al., 2017; Cai et al., 2018). Interestingly, no similar cis-elements distribution was shown in promoter regions of BlSPLs within the same phylogenetic group (Figure 6), implying the function divergences of SPL genes was not only reflected in the coding regions, but also occurred in the promoter regions.

In order to further explore the potential roles of BlSPL genes in growth and development of B. luminifera, the expression profiles of BlSPL genes were analyzed in 16 different tissues/organs. Considering the correlation between BlSPLs and miR156, the expression profile of mature BlmiR156 was also analyzed accordingly. Different expression patterns of the BlSPL genes were revealed (Figure 7). A few BlSPL genes (BlSPL4, BlSPL5, BlSPL10, and BlSPL12) presented a constitutive expression pattern in all tissues or organs examined, the remaining majority exhibited a development- and tissue-dependent expression patterns. On the other hand, it is worthy to note that the differentiated expression pattern was usually observed in the BlmiR156-targeted BlSPL genes, while more than half of the BlmiR156-nontargeted BlSPL genes tended to exhibit the constitutive expression pattern. Similar expression differences between miR156-targeted and nontargeted SPL genes have been revealed in apple, soybean, Brassica napus and S.miltiorrhiza (Li et al., 2013; Zhang et al., 2013; Cheng et al., 2016; Tripathi et al., 2017). This means that miR156-targeted and nontargeted SPL genes experienced divergent evolution with differentiated expression patterns, suggesting their distinct roles in development and growth of plants. Moreover, the BlmiR156 showed a differentiated expression pattern in different tissues/organs, with highest expression level in seeds and comparative low level in most of the remaining tissues/organs (Figure 7). Notably, most of the BlmiR156-targeted SPL genes exhibited rather low expressions in seeds. This is consistent with the results from apple, in which miR156 was also highly expressed in seeds of young and mature fruit, and many miR156-targeted MdSPL genes showed extremely low expression levels in seeds (Li et al., 2013). If taken the seeds as the plants with minimum physiological age, it implies the tight correlation of miR156/SPL module with vegetative phase changes in trees.

Because of the important regulatory roles of miR156/SPL module in vegetative phase change, the expressions of BlSPLs and BlmiR156 were evaluated in seedlings/plants with different physiological ages. As expected, most BlmiR156-targeted BlSPL genes were dramatically up-regulated with ages, whereas most BlmiR156-nontargeted BlSPL genes tend to be constitutively expressed at different ages (Figure 8). Meanwhile, the expression of BlmiR156 was obviously decreased with ages (Figure 8), which is correlated negatively with the expression patterns of those BlmiR156-targeted BlSPL genes. Such age-dependent expression patterns of miR156 and miR156-targeted SPL genes have also been reported in Arabidopsis, Arabis alpine and Populus (Wang et al., 2011; Bergonzi et al., 2013; Xu et al., 2016). Further functional studies have indicated the important roles of miR156/SPL module in phase transition. In Arabidopsis, constitutive expression of miR156 causes prolonged expression of juvenile vegetative traits and delayed flowering (Wu and Poethig, 2006). AtSPL3/4/5 directly activate LEAFY (LFY), FRUITFULL (FUL), and APETALA1 (AP1) to redundantly promote flowering in Arabidopsis (Yamaguchi et al., 2009). In addition, AtSPL9 and AtSPL10 also act redundantly to promote the transcription of miR172, which further promotes the transition from the juvenile to the adult phase (Wu et al., 2009). Similar functions of miR156 and SPL genes have also been revealed in Chinese cabbage (Wang et al., 2014), rice (Xie et al., 2006) and perennial plants such as the tree P x Canadensis (Wang et al., 2011), Cardamine flexuosa (Zhou et al., 2013). These evidences show that the miR156/SPL module plays a conserved role in regulating vegetative phase change in diverse plant species. Thus, these miR156-targeted BlSPLs must also have important functions during vegetative phase change of B. luminifera, and could be considered as molecular markers for reaching to reproductive phase. Additionally, it was also noticed that four BlmiR156-targeted BlSPL genes (BlSPL15-BlSPL18) showed their highest expression levels at 1.5-year-old, when the B. luminifera plants start to flowering. While, the other BlmiR156-targeted BlSPLs (except for BlSPL11) reached to their highest expression level at 4-year-old (Figure 8). This expression difference suggested that these miR156-targeted BlSPL genes may not only have overlapping functions, but also play important roles in different stages of phase transition.

DELLA proteins, act as central transcriptional repressors of GA responses, have been found to interact and repress SPL proteins activities. The physical interaction between SPL and DELLA is thought to be an integrator of the age and GA pathways in flowering (Yu et al., 2012). The widespread interactions between DELLAs and miR156-targeted SPLs have been confirmed in Arabidopsis by yeast two-hybrid. For instance, RGA could interact with AtSPL2, AtSPL9 and AtSPL11, and RGL1 interacts with AtSPL2, AtSPL10, and AtSPL11 (Yu et al., 2012). Consistently, the interactions between BlRGA and miR156-targeted BlSPLs (BlSPL1/6/8/13/18), as well as BlRGL and miR156-targeted BlSPLs (BlSPL1/8/13/16/18), were also detected in this study (Figure 9). However, not all miR156-targeted BlSPLs could interact with these two DELLA proteins, and the interaction strengths were apparently varied in different combinations (Supplementary Figure S3). It means the interactions between DELLA proteins and SPLs are conserved regulation mechanisms for flowering induction in plants, but not all miR156-targeted SPL genes take part in this process. More interestingly, two miR156-nontargeted BlSPLs (BlSPL2 and BlSPL5) were also revealed to interact with these two DELLA proteins, which were reported for the first time in our study. This implies that interactions between DELLA proteins and miR156-nontargeted SPLs may exert important functions in other GA regulated processes.

Conclusion

This study here represents the first genome-wide characterization of SPL genes in B. luminifera. Comprehensive analyses were performed to characterize 18 predicted BlSPLs, including their sequence features, phylogeny, intron/exon structure, conserved motif, miR156-mediated posttranscriptional regulation, cis-elements in promoter regions and expression patterns. Furthermore, the interactions between BlSPLs and DELLA proteins were also investigated to explore their roles in GA regulated biological processes. The results showed that 18 BlSPLs were clustered into 9 groups, and most BlSPLs of the same group exhibit high conservation in sequence features, intron/exon structures, motif composition, and posttranscriptional regulation. Moreover, two distinct expression patterns were revealed for miR156-targeted and nontargeted BlSPL genes in different tissues/organs, suggesting the differentiated roles of these BlSPLs in development and growth of B. luminifera. The expression analysis in plants with different ages showed that miR156/SPL module may also play important roles in regulating vegetative phase change in B. luminifera. In addition, protein interaction assay indicated that BlSPL genes participate in GA regulated biological processes through physical interacting with DELLA proteins. These results provide useful information that facilitates further elucidation of the potential biological roles of BlSPL genes in B. luminifera.

Author Contributions

E-PL, X-YL, and Z-KT: conceived and designed the experiments; E-PL, X-YL, and M-YN: performed the experiments; H-HH, Z-KT, and J-HZ: contributed reagents, materials, analysis tools; E-PL and X-YL: analyzed the data; E-PL and X-YL: wrote the paper.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

This work was supported by the Zhejiang Provincial Natural Science Foundation of China (No. LY15C160008), the National Natural Science foundation of China (No. 31770641, 31300566 and 31470674) and the Tree Breeding Research Project of Zhejiang Province, China (No. 2016C02056-2).

Supplementary Material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fpls.2018.00608/full#supplementary-material

Supplementary Figure S1. Autoactivation and toxicity test for the BD fused constructs and AD fused constructs. (A) Autoactivation and toxicity of BD-BlSPLs, and AD-BlRGA and AD-BlRGL. (B) Autoactivation and toxicity of BD- BlRGA, BD-BlRGL, and AD-BlSPL6/7/10. BD, GAL4 binding domain; AD, GAL4 activation domain. Autoactivation was detected on the SD-Leu-Trp-His-Ade plates supplemented with 40 μg/ml X-α-gal and 125 ng/ml Aureobasidin A (SD-L-W-H-A-X-AbA). Toxicity was examined on the SD medium without leucine and tryptophan (SD-L-W). The yeast clones were grown with dilutions to 10−1 and 10−2.

Supplementary Figure S2. Alignment of ANK/ANK-2 domain in BlSPL4, BlSPL5 and BlSPL10. The ANK/ANK-2 domain is marked by solid lines.

Supplementary Figure S3. The interaction strengths were determined through β-galactosidase activity assay. (A) The interaction strengths between BlSPLs and BlRGA. (B) The interaction strengths between BlSPLs and BlRGL. Empty, BD and AD vector without target genes, as a negative control.

Supplementary Figure S4. Expression profiles of BlRGA and BlRGL. (A) Expression patterns of BlRGA in different tissues/organs. (B) Expression patterns of BlRGL in different tissues/organs. (C) Expression patterns of BlRGA in plants with different ages. (D) Expression patterns of BlRGL in plants with different ages. The same samples as figure 7 and 8 were used for the expression profiling of BlRGA and BlRGL. Transcript levels in buds and 5-day-old (5 d) seedlings were arbitrarily set to 1 and the levels in other tissues were given relative to this. Error bars represent standard deviations of mean value from three biological replicates and four technical replicates.

Supplementary Table S1. The mapping coordinates of TBLASTN by querying AtSPLs and PtSPLs protein sequences against draft genome of B. luminifera.

Supplementary Table S2. The predicted and corrected gene models of BlSPL genes.

Supplementary Table S3. Primer names and their sequences used for gene cloning.

Supplementary Table S4. Primer names and their sequences used for 5' RLM-RACE.

Supplementary Table S5. Primer names and their sequences used for qRT-PCR.

Supplementary Table S6. Primer names and their sequences used for ORF cloning.

Supplementary Table S7. General information of SPL genes used for phylogenetic analysis.

Supplementary Table S8. Data sources of SPL genes used for phylogenetic analysis.

Supplementary Table S9. Cis-elements in the promoter region of 18 BlSPL genes.

Supplementary Table S10. The pearson's correlation coefficients between expressions of two DELLA and BlSPL genes.

References

Altschul, S. F., Madden, T. L., Schäffer, A. A., Zhang, J., Zhang, Z., Miller, W., et al. (1997). Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 25, 3389–3402. doi: 10.1093/nar/25.17.3389

PubMed Abstract | CrossRef Full Text | Google Scholar

Bailey, T. L., and Elkan, C. (1994). Fitting a mixture model by expectation maximization to discover motifs in biopolymers. Proc. Int. Conf. Intell. Syst. Mol. Biol. 2, 28–36.

PubMed Abstract | Google Scholar

Bergonzi, S., Albani, M. C., Ver, L. V. T. E., Nordström, K. J., Wang, R., Schneeberger, K., et al. (2013). Mechanisms of age-dependent response to winter temperature in perennial flowering of Arabis alpina. Science 340, 1094–1097. doi: 10.1126/science.1234116

PubMed Abstract | CrossRef Full Text | Google Scholar

Birkenbihl, R. P., Jach, G., Saedler, H., and Huijser, P. (2005). Functional dissection of the plant-specific SBP-domain: overlap of the DNA-binding and nuclear localization domains. J. Mol. Biol. 352, 585–596. doi: 10.1016/j.jmb.2005.07.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Cai, C., Guo, W., and Zhang, B. (2018). Genome-wide identification and characterization of SPL transcription factor family and their evolution and expression profiling analysis in cotton. Sci. Rep. 8:762. doi: 10.1038/s41598-017-18673-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Cardon, G., Höhmann, S., Klein, J., Nettesheim, K., Saedler, H., and Huijser, P. (1999). Molecular characterisation of the Arabidopsis SBP-box genes. Gene 237, 91–104. doi: 10.1016/S0378-1119(99)00308-X

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, X. (2009). Small RNAs and their roles in plant development. Annu. Rev. Cell Dev. Biol. 25, 21–44. doi: 10.1146/annurev.cellbio.042308.113417

PubMed Abstract | CrossRef Full Text | Google Scholar

Cheng, H., Hao, M., Wang, W., Mei, D., Tong, C., Wang, H., et al. (2016). Genomic identification, characterization and differential expression analysis of SBP-box gene family in Brassica napus. BMC Plant Biol. 16:196. doi: 10.1186/s12870-016-0852-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Cui, L. G., Shan, J. X., Shi, M., Gao, J. P., and Lin, H. X. (2014). The miR156-SPL9-DFR pathway coordinates the relationship between development and abiotic stress tolerance in plants. Plant J. 80, 1108–1117. doi: 10.1111/tpj.12712

PubMed Abstract | CrossRef Full Text | Google Scholar

Dai, X., and Zhao, P. X. (2011). psRNATarget: a plant small RNA target analysis server. Nucleic Acids Res. 39, W155–W159. doi: 10.1093/nar/gkr319

PubMed Abstract | CrossRef Full Text | Google Scholar

Davière, J. M., Wild, M., Regnault, T., Baumberger, N., Eisler, H., Genschik, P., et al. (2014). Class I TCP-DELLA interactions in inflorescence shoot apex determine plant height. Curr. Biol. 24, 1923–1928. doi: 10.1016/j.cub.2014.07.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Gandikota, M., Birkenbihl, R. P., Höhmann, S., Cardon, G. H., Saedler, H., and Huijser, P. (2007). The miRNA156/157 recognition element in the 3' UTR of the Arabidopsis SBP box gene SPL3 prevents early flowering by translational inhibition in seedlings. Plant J. 49, 683–693. doi: 10.1111/j.1365-313X.2006.02983.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Gou, J., Fu, C., Liu, S., Tang, C., Debnath, S., Flanagan, A., et al. (2017). The miR156-SPL4 module predominantly regulates aerial axillary bud formation and controls shoot architecture. New Phytol. 216, 829–840. doi: 10.1111/nph.14758

PubMed Abstract | CrossRef Full Text | Google Scholar

Gou, J. Y., Felippes, F. F., Liu, C. J., Weigel, D., and Wang, J. W. (2011). Negative regulation of anthocyanin biosynthesis in Arabidopsis by a miR156-targeted SPL transcription factor. Plant Cell 23, 1512–1522. doi: 10.1105/tpc.111.084525

PubMed Abstract | CrossRef Full Text | Google Scholar

Guo, A. Y., Zhu, Q. H., Gu, X., Ge, S., Yang, J., and Luo, J. (2008). Genome-wide identification and evolutionary analysis of the plant specific SBP-box transcription factor family. Gene 418, 1–8. doi: 10.1016/j.gene.2008.03.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Guo, H. S., Xie, Q., Fei, J. F., and Chua, N. H. (2005). MicroRNA directs mRNA cleavage of the transcription factor NAC1 to downregulate auxin signals for arabidopsis lateral root development. Plant Cell 17, 1376–1386. doi: 10.1105/tpc.105.030841

PubMed Abstract | CrossRef Full Text | Google Scholar

Hu, B., Jin, J., Guo, A. Y., Zhang, H., Luo, J., and Gao, G. (2015). GSDS 2.0: an upgraded gene feature visualization server. Bioinformatics 31, 1296–1297. doi: 10.1093/bioinformatics/btu817

PubMed Abstract | CrossRef Full Text | Google Scholar

Huang, D., Wang, S., Zhang, B., Shang-Guan, K., Shi, Y., Zhang, D., et al. (2015). A gibberellin-mediated DELLA-NAC signaling cascade regulates cellulose synthesis in rice. Plant Cell 27, 1681–1696. doi: 10.1105/tpc.15.00015

PubMed Abstract | CrossRef Full Text | Google Scholar

Jiao, Y., Wang, Y., Xue, D., Wang, J., Yan, M., Liu, G., et al. (2010). Regulation of OsSPL14 by OsmiR156 defines ideal plant architecture in rice. Nat. Genet. 42, 541–544. doi: 10.1038/ng.591

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, J. J., Lee, J. H., Kim, W., Jung, H. S., Huijser, P., and Ahn, J. H. (2012). The microRNA156-SQUAMOSA PROMOTER BINDING PROTEIN-LIKE3 module regulates ambient temperature-responsive flowering via FLOWERING LOCUS T in Arabidopsis. Plant Physiol. 159, 461–478. doi: 10.1104/pp.111.192369

PubMed Abstract | CrossRef Full Text | Google Scholar

Klein, J., Saedler, H., and Huijser, P. (1996). A new family of DNA binding proteins includes putative transcriptional regulators of the Antirrhinum majus floral meristem identity gene SQUAMOSA. Mol. Gen. Genet. 250, 7–16.

PubMed Abstract | Google Scholar

Kumar, S., Stecher, G., and 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

PubMed Abstract | CrossRef Full Text | Google Scholar

Lal, S., Pacis, L. B., and Smith, H. M. (2011). Regulation of the SQUAMOSA PROMOTER-BINDING PROTEIN-LIKE genes/microRNA156 module by the homeodomain proteins PENNYWISE and POUND-FOOLISH in Arabidopsis. Mol. Plant 4, 1123–1132. doi: 10.1093/mp/ssr041

PubMed Abstract | CrossRef Full Text | Google Scholar

Lescot, M., Déhais, P., Thijs, G., Marchal, K., Moreau, Y., Van de Peer, Y., et al. (2002). PlantCARE, a database of plant cis-acting regulatory elements and a portal to tools for in silico analysis of promoter sequences. Nucleic Acids Res. 30, 325–327. doi: 10.1093/nar/30.1.325

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, C., and Lu, S. (2014). Molecular characterization of the SPL gene family in Populus trichocarpa. BMC Plant Biol. 14:131. doi: 10.1186/1471-2229-14-131

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, J., Hou, H., Li, X., Xiang, J., Yin, X., Gao, H., et al. (2013). Genome-wide identification and analysis of the SBP-box family genes in apple (Malus x domestica Borkh.). Plant Physiol. Biochem. 70, 100–114. doi: 10.1016/j.plaphy.2013.05.021

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu, W. Z., Niu, M. Y., Li, X. Y., Lin, E. P., Huang, H. H., and Tong, Z. K. (2016). The selection of reference genes for quantitative PCR in Betula luminifera. Sci. Silva. Sin. 52, 29–37. doi: 10.11707/j.1001-7488.20160804

CrossRef Full Text

Locascio, A., Blazquez, M. A., and Alabadi, D. (2013). Genomic analysis of DELLA protein activity. Plant Cell Physiol. 54, 1229–1237. doi: 10.1093/pcp/pct082

PubMed Abstract | CrossRef Full Text | Google Scholar

Michaely, P., and Bennett, V. (1992). The ANK repeat: a ubiquitous motif involved in macromolecular recognition. Trends Cell Biol. 2, 127–129. doi: 10.1016/0962-8924(92)90084-Z

PubMed Abstract | CrossRef Full Text | Google Scholar

Miura, K., Ikeda, M., Matsubara, A., Song, X. J., Ito, M., Asano, K., et al. (2010). OsSPL14 promotes panicle branching and higher grain productivity in rice. Nat. Genet. 42, 545–549. doi: 10.1038/ng.592

PubMed Abstract | CrossRef Full Text | Google Scholar

Pan, F., Wang, Y., Liu, H., Wu, M., Chu, W., Chen, D., et al. (2017). Genome-wide identification and expression analysis of SBP-like transcription factor genes in Moso Bamboo (Phyllostachys edulis). BMC Genomics 18:486. doi: 10.1186/s12864-017-3882-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Rhoades, M. W., Reinhart, B. J., Lim, L. P., Burge, C. B., Bartel, B., and Bartel, D. P. (2002). Prediction of plant microRNA targets. Cell 110, 513–520. doi: 10.1016/S0092-8674(02)00863-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Riese, M., Hohmann, S., Saedler, H., Munster, T., and Huijser, P. (2007). Comparative analysis of the SBP-box gene families in P. patens and seed plants. Gene 401, 28–37. doi: 10.1016/j.gene.2007.06.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Rombolá-Caldentey, B., Rueda-Romero, P., Iglesias-Fernández, R., Carbonero, P., and Oñate-Sánchez, L. (2014). Arabidopsis DELLA and Two HD-ZIP transcription factors regulate GA signaling in the epidermis through the L1 Box cis-Element. Plant Cell 26, 2905–2919. doi: 10.1105/tpc.114.127647

PubMed Abstract | CrossRef Full Text | Google Scholar

Rossignol, P., Collier, S., Bush, M., Shaw, P., and Doonan, J. H. (2007). Arabidopsis POT1A interacts with TERT-V(I8), an N-terminal splicing variant of telomerase. J Cell Sci. 120(Pt 20), 3678–3687. doi: 10.1242/jcs.004119

PubMed Abstract | CrossRef Full Text | Google Scholar

Salinas, M., Xing, S., Höhmann, S., Berndtgen, R., and Huijser, P. (2011). Genomic organization, phylogenetic comparison and differential expression of the SBP-box family of transcription factors in tomato. Planta 235, 1171–1184. doi: 10.1007/s00425-011-1565-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Schmittgen, T. D., and Livak, K. J. (2008). Analyzing real-time PCR data by the comparative C(T) method. Nat. Protoc. 3, 1101–1108. doi: 10.1038/nprot.2008.73

PubMed Abstract | CrossRef Full Text | Google Scholar

Schwarz, S., Grande, A. V., Bujdoso, N., Saedler, H., and Huijser, P. (2008). The microRNA regulated SBP-box genes SPL9 and SPL15 control shoot maturation in Arabidopsis. Plant Mol. Biol. 67, 183–195. doi: 10.1007/s11103-008-9310-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Shalom, L., Shlizerman, L., Zur, N., Doron-Faigenboim, A., Blumwald, E., and Sadka, A. (2015). Molecular characterization of SQUAMOSA PROMOTER BINDING PROTEIN-LIKE (SPL) gene family from Citrus and the effect of fruit load on their expression. Front. Plant Sci. 6:389. doi: 10.3389/fpls.2015.00389

PubMed Abstract | CrossRef Full Text | Google Scholar

Shao, F., Lu, Q., Wilson, I. W., and Qiu, D. (2017). Genome-wide identification and characterization of the SPL gene family in Ziziphus jujuba. Gene 627, 315–321. doi: 10.1016/j.gene.2017.06.044

PubMed Abstract | CrossRef Full Text | Google Scholar

Solovyev, V., Kosarev, P., Seledsov, I., and Vorobyev, D. (2006). Automatic annotation of eukaryotic genes, pseudogenes and promoters. Genome Biol. 7(Suppl 1), S10–S11. doi: 10.1186/gb-2006-7-s1-s10

PubMed Abstract | CrossRef Full Text | Google Scholar

Song, A., Gao, T., Wu, D., Xin, J., Chen, S., Guan, Z., et al. (2016). Transcriptome-wide identification and expression analysis of chrysanthemum SBP-like transcription factors. Plant Physiol. Biochem. 102, 10–16. doi: 10.1016/j.plaphy.2016.02.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Sun, T. P. (2010). Gibberellin-GID1-DELLA: a pivotal regulatory module for plant growth and development. Plant Physiol. 154, 567–570. doi: 10.1104/pp.110.161554

PubMed Abstract | CrossRef Full Text | Google Scholar

Tripathi, R. K., Goel, R., Kumari, S., and Dahuja, A. (2017). Genomic organization, phylogenetic comparison, and expression profiles of the SPL family genes and their regulation in soybean. Dev. Genes Evol. 227, 101–119. doi: 10.1007/s00427-017-0574-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Varkonyi-Gasic, E., Wu, R., Wood, M., Walton, E., and Hellens, R. (2007). Protocol: a highly sensitive RT-PCR method for detection and quantification of microRNAs. Plant Methods 3:12. doi: 10.1186/1746-4811-3-12

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, H., and Wang, H. (2015). The miR156/SPL module, a regulatory hub and versatile toolbox, gears up crops for enhanced agronomic traits. Mol. Plant 8, 677–688. doi: 10.1016/j.molp.2015.01.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, J. W., Czech, B., and Weigel, D. (2009). miR156-regulated SPL transcription factors define an endogenous flowering pathway in Arabidopsis thaliana. Cell 138, 738–749. doi: 10.1016/j.cell.2009.06.014

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, J. W., Park, M. Y., Wang, L. J., Koo, Y., Chen, X. Y., Weigel, D., et al. (2011). miRNA control of vegetative phase change in trees. PLoS Genet. 7:e1002012. doi: 10.1371/journal.pgen.1002012

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, L., Liu, N., Wang, T., Li, J., Wen, T., Yang, X., et al. (2017). The GhmiR157a/GhSPL10 regulatory module controls initial cellular dedifferentiation and callus proliferation in cotton by modulating ethylene-mediated flavonoid biosynthesis. J. Exp. Bot. 69, 1081–1093. doi: 10.1093/jxb/erx475

CrossRef Full Text | Google Scholar

Wang, Y., Wu, F., Bai, J., and He, Y. (2014). BrpSPL9 (Brassica rapa ssp. pekinensis SPL9) controls the earliness of heading time in Chinese cabbage. Plant Biotechnol. J. 12, 312–321. doi: 10.1111/pbi.12138

PubMed Abstract | CrossRef Full Text | Google Scholar

Willems, E., Leyns, L., and Vandesompele, J. (2008). Standardization of real-time PCR gene expression data from independent biological replicates. Anal. Biochem. 379, 127–129. doi: 10.1016/j.ab.2008.04.036

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, G., Park, M. Y., Conway, S. R., Wang, J. W., Weigel, D., and Poethig, R. S. (2009). The sequential action of miR156 and miR172 regulates developmental timing in Arabidopsis. Cell 138, 750–759. doi: 10.1016/j.cell.2009.06.031

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, G., and Poethig, R. S. (2006). Temporal regulation of shoot development in Arabidopsis thaliana by miR156 and its target SPL3. Development 133, 3539–3547. doi: 10.1242/dev.02521

PubMed Abstract | CrossRef Full Text | Google Scholar

Xie, K., Wu, C., and Xiong, L. (2006). Genomic organization, differential expression, and interaction of SQUAMOSA promoter-binding-like transcription factors and microRNA156 in rice. Plant Physiol. 142, 280–293. doi: 10.1104/pp.106.084475

PubMed Abstract | CrossRef Full Text | Google Scholar

Xu, M., Hu, T., Zhao, J., Park, M. Y., Earley, K. W., Wu, G., et al. (2016). Developmental Functions of miR156-Regulated SQUAMOSA PROMOTER BINDING PROTEIN-LIKE (SPL) Genes in Arabidopsis thaliana. PLoS Genet. 12:e1006263. doi: 10.1371/journal.pgen.1006263

PubMed Abstract | CrossRef Full Text | Google Scholar

Yamaguchi, A., Wu, M. F., Yang, L., Wu, G., Poethig, R. S., and Wagner, D. (2009). The microRNA-regulated SBP-Box transcription factor SPL3 is a direct upstream activator of LEAFY, FRUITFULL, and APETALA1. Dev. Cell 17, 268–278. doi: 10.1016/j.devcel.2009.06.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Yamaguchi, N., Winter, C. M., Wu, M. F., Kanno, Y., Yamaguchi, A., Seo, M., et al. (2014). Gibberellin acts positively then negatively to control onset of flower formation in Arabidopsis. Science 344, 638–641. doi: 10.1126/science.1250498

PubMed Abstract | CrossRef Full Text | Google Scholar

Yamasaki, K., Kigawa, T., Inoue, M., Tateno, M., Yamasaki, T., Yabuki, T., et al. (2004). A novel zinc-binding motif revealed by solution structures of DNA-binding domains of Arabidopsis SBP-family transcription factors. J. Mol. Biol. 337, 49–63. doi: 10.1016/j.jmb.2004.01.015

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang, Z., Wang, X., Gu, S., Hu, Z., Xu, H., and Xu, C. (2008). Comparative study of SBP-box gene family in Arabidopsis and rice. Gene 407, 1–11. doi: 10.1016/j.gene.2007.02.034

PubMed Abstract | CrossRef Full Text | Google Scholar

Yu, D., Hu, Y., Wang, H., Pan, J., Li, Y., and Lou, D. (2016). The DELLA-CONSTANS transcription factor cascade integrates gibberelic acid and photoperiod signaling to regulate flowering. Plant Physiol. 172, 479–488. doi: 10.1104/pp.16.00891

PubMed Abstract | CrossRef Full Text | Google Scholar

Yu, S., Galvão, V. C., Zhang, Y. C., Horrer, D., Zhang, T. Q., Hao, Y. H., et al. (2012). Gibberellin regulates the Arabidopsis floral transition through miR156-targeted SQUAMOSA promoter binding-like transcription factors. Plant Cell 24, 3320–3332. doi: 10.1105/tpc.112.101014

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, B., Pan, X., Cobb, G. P., and Anderson, T. A. (2006). Plant microRNA: a small regulatory molecule with big impact. Dev. Biol. 289, 3–16. doi: 10.1016/j.ydbio.2005.10.036

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, H. X., Jin, J. H., He, Y. M., Lu, B. Y., Li, D. W., Chai, W. G., et al. (2016). Genome-Wide Identification and Analysis of the SBP-Box Family Genes under Phytophthora capsici Stress in Pepper (Capsicum annuum L.). Front. Plant Sci. 7:504. doi: 10.3389/fpls.2016.00504

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, L., Wu, B., Zhao, D., Li, C., Shao, F., and Lu, S. (2013). Genome-wide analysis and molecular dissection of the SPL gene family in Salvia miltiorrhiza. J. Integr. Plant Biol. 56, 38–50. doi: 10.1111/jipb.12111

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, Y., Schwarz, S., Saedler, H., and Huijser, P. (2007). SPL8, a local regulator in a subset of gibberellin-mediated developmental processes in Arabidopsis. Plant Mol. Biol. 63, 429–439. doi: 10.1007/s11103-006-9099-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhou, C. M., Zhang, T. Q., Wang, X., Yu, S., Lian, H., Tang, H., et al. (2013). Molecular basis of age-dependent vernalization in Cardamine flexuosa. Science 340, 1097–1100. doi: 10.1126/science.1234340

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: Betula luminifera, SPL gene family, miR156, expression patterns, vegetative phase change, DELLA proteins

Citation: Li X-Y, Lin E-P, Huang H-H, Niu M-Y, Tong Z-K and Zhang J-H (2018) Molecular Characterization of SQUAMOSA PROMOTER BINDING PROTEIN-LIKE (SPL) Gene Family in Betula luminifera. Front. Plant Sci. 9:608. doi: 10.3389/fpls.2018.00608

Received: 05 February 2018; Accepted: 17 April 2018;
Published: 04 May 2018.

Edited by:

Rosalba Giugno, University of Verona, Italy

Reviewed by:

Gaurav Sablok, University of Helsinki, Finland
Valentino Ruggieri, CONSORCIO CSIC-IRTA-UAB-UB Centro de Investigación Agrigenómica (CRAG), Spain

Copyright © 2018 Li, Lin, Huang, Niu, Tong and Zhang. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Er-Pei Lin, zjulep@hotmail.com
Zai-Kang Tong, zktong@zafu.edu.cn

These authors have contributed equally to this work.

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.