Skip to main content

Genome-wide identification of sugar transporter gene family in Brassicaceae crops and an expression analysis in the radish

Abstract

Background

Sugar not only is an important biomacromolecule that plays important roles in plant growth, development, and biotic and abiotic stress tolerance but also provides a skeleton for other macromolecules, such as proteins and nucleic acids. Sugar transporter proteins (STPs) play essential roles in plant sugar transport and ultimately affect the abovementioned life processes. However, the evolutionary dynamics of this important gene family in Brassicaceae crops are still largely unknown, and the functional differentiation of radish STP genes remains unclear.

Results

In the present study, a comparative genomic study of STP genes in five representative Brassicaceae crops was conducted, and a total of 25, 25, 28, 36 and 49 STP genes were individually identified in Raphanus sativus (Rs), Brassica oleracea (Bo), B. rapa (Br), B. napus (Bn) and B. juncea (Bj), which were divided into four clades by phylogenetic analysis. The number of STP genes was no direct correlation with genome size and the total number of coding genes in Brassicaceae crops, and their physical and chemical properties showed no significant difference. Expression analysis showed that radish STP genes play vital roles not only in flower and seedpod development but also under heavy metal (cadmium, chromium and lead), NaCl and PEG-6000 stresses, Agrobacterium tumefaciens infection, and exogenous sugar treatment. RsSTP13.2 was significantly upregulated in the resistant radish cultivar by A. tumefaciens infection and induced by heavy metal, NaCl and PEG-6000 stress, indicating that it is involved in resistance to both biotic and abiotic stress in radish.

Conclusions

The present study provides insights into the evolutionary patterns of the STP gene family in Brassicaceae genomes and provides a theoretical basis for future functional analysis of STP genes in Brassicaceae crops.

Peer Review reports

Background

Most plants on earth, except for parasitic plants, fix carbon through photosynthesis to produce soluble sugar as the main carbohydrate through various reactions in the cytoplasm [1]. These sugars have the functions of providing energy for cell life activities, providing a skeleton for macromolecules such as proteins and nucleic acids. Sugars also participate in the regulation of various metabolic pathways and biotic and abiotic stress responses of plants as signal molecules [2, 3].

Sugars synthesized from source cells were transported to the sink organs via the phloem mainly in the form of sucrose [4]. In the sink organs, sucrose can directly enter the cells by symplast pathways through plasmodesmata or via the apoplast pathway mediated by sugar transporters [1, 4]. In addition, sucrose can also be hydrolysed to monosaccharides glucose and fructose in the apoplast by cell wall invertases, and thereafter, was transported into the cells across the plasma membrane mediated by monosaccharide transporters [4]. At present, three types of sugar transporters have been found in plants: sugars will eventually be exported transporters (SWEETs), monosaccharide transporters (MSTs) and sucrose transporters (SUTs) [5]. MSTs can be further divided into seven subfamilies, including sugar transporter protein (STP), vacuolar glucose transporter (VGT), tonoplast monosaccharide transporter (TMT), plastidic glucose transporter (GlcT)/suppressor of G protein beta 1 (SGB1), polyol transporter (PLT), inositol transporter (INT) and early response to dehydration-6-like (ERD6L) [6]. Among these subfamilies, the STPs are the best characterized subfamily and function in transporting hexose from the apoplastic space into the cell [6].

STPs contain 12 transmembrane domains and are located on plasma membranes, regarded as H+/sugar symporters [7]. There are 14 members (AtSTP1-AtSTP14) of the STP gene family in the Arabidopsis genome [8], and AtSTP1 is the first sugar transporter cloned from a higher plant [9]. To date, all 14 AtSTPs have been studied by heterologous expression in yeast, and most of them have a broad spectrum of substrates. AtSTP1 [9], AtSTP2 [10], AtSTP4 [11], AtSTP6 [12], AtSTP11 [13] and AtSTP13 [4, 14] were proven to transporter hexose glucose, galactose and mannose, pentose xylose and nonmetabolisable 3-O-methylglucose, while AtSTP9 is a glucose-specific transporter [15]. AtSTP7 cannot transport the abovementioned monosaccharides but is specific for the pentoses l-arabinose and d-xylose [6, 16], and AtSTP3 [8] has uptake activity for the abovementioned monosaccharides except for fructose. AtSTP5 might be a pseudogne [17]. AtSTP10 encodes a protein that catalyses the uptake of glucose, galactose and mannose but not fructose [18]. AtSTP8 was reported to have broad-spectrum monosaccharides such as glucose, fructose, mannose, galactose, arabinose and xylose [16, 17]. AtSTP12 can transport glucose, galactose and mannose [16], while AtSTP14 does not take up glucose and fructose but can transport galactose [19].

Previous studies have indicated that STPs are involved in regulating multiple growth and developmental processes and biotic and abiotic resistance by regulating the distribution and accumulation of soluble sugars in plants. Overexpression of AtSTP1 in A. thaliana significantly alters the extracellular sugar contents and inhibits its growth and branching [20]. AtSTP1 and AtSTP4 cooperate to import glucose to guard cells, providing carbon sources for light-induced stomatal opening and guarding cell starch accumulation [21]. AtSTP4/6/8/9/10/11 displays high expression levels in pollen tubes, and a sextuple knockout plant eliminates the inhibitory effect of glucose on pollen tube elongation in vitro [22]. The expression of AtSTP13 was significantly induced by Botrytis cinerea infection, and its overexpression enhanced the resistance to grey mould disease by improving glucose uptake [23], while its orthologous gene in wheat, TaSTP13, contributes to susceptibility to Puccinia striiformis f. sp. tritici (Pst), most likely by increasing the fungal sugar supply [24]. In addition, TaSTP6 was induced in leaves by Pst infection, and its expression promoted the susceptibility of wheat to stripe rust [25]. A more recent study indicated that overexpression of AtSTP8 in A. thaliana significantly induced the accumulation of hexose in the mature leaf and enhanced susceptibility to powdery mildew disease [17]. The expression of STPs is also significantly affected by abiotic stress. For example, both excess zinc and iron/zinc deficient stress significantly upregulate the expression of STP13 in shoot of Phaseolus vulgaris L., while excess zinc stress upregulated and iron/zinc deficient stress downregulated its expression in roots [26]. The expression of STP2, STP3, STP4, STP11, STP19 and STP25 in rice was significantly induced under salt, osmotic, and drought stress. STP11 was upregulated under ABA, IAA, 6-BA, SA and GA treatment, while STP1 and STP14 were upregulated under sucrose, glucose and fructose treatment [27]. These findings indicate that STPs play important roles in plant sugar transport, growth, development, and stress tolerance.

Brassicaceae includes many economically and nutritionally important crops, such as Raphanus sativus, Brassica oleracea, B. rapa, B. napus and B. juncea. The ancestor of diploid Brassica and Raphanus species, including B. oleracea, B. rapa, and R. sativus, has undergone Brassiceae lineage-specific whole-genome triplication (WGT) after its divergence from the A. thaliana lineage approximately 20 million years ago (MYA) [28]. However, the neotetraploids B. napus and B. juncea were allopolyploid between the ancestors of B. rapa (genome AA) and B. oleracea (genome CC), B. rapa (genome AA) and Brassica nigra (genome BB) [28]. The expansion and contraction of the STP gene family in these plants during their evolution are still largely unknown, even though high-quality genomes have been assembled. Thus, a genome-wide identification and analysis of the STP gene family in R. sativus, B. oleracea, B. rapa, B. napus and B. juncea was conducted in the present study. Additionally, the expression patterns in various organs and in response to biotic and abiotic stress were determined in the radish to characterize the functional differentiation of the STP gene family. The results of the present study provide insight into the expansion and contraction of the STP gene family in Brassicaceae and provide the basis for their functional study.

Results

Genome-wide identification of STP genes from five Brassicaceae crops

Twenty-five, 25, 28, 36 and 49 STP-encoding genes were identified from the genomes of R. sativus (Rs), B. oleracea (Bo), B. rapa (Br), B. napus (Bn) and B. juncea (Bj), respectively (Table S1). There were also 14 A. thaliana STP genes identified in this study, which were identical to those described in a previous report [4], confirming the reliability of our results. No STP gene was found on the scaffolds of these six genomes, while the number of STP genes distributed on chromosomes (Chr) of each species varied greatly. In R. sativus, Chr1 and Chr5 had the highest number of STP genes. B. oleracea and B. rapa had the most STP genes on chromosomes C1 and A01, respectively. B. napus An-subgenomes had the most genes on A03, and Cn-subgenomes had the most genes on C05 and C07. The B. juncea An and Bn subgenomes had the most genes on A01 and B01, respectively (Fig. S1).

The physical and chemical properties of all STP proteins were analysed (Table S2). There were no significant differences in amino acid residue number, molecular weights, aromaticity, instability index, isoelectric point or gravy among the six species (Fig. 1). The predicted aromaticity ranged from 0.09–0.15, the instability index ranged from 28.58–45.46, and the isoelectric point ranged from 5.44–9.65 (Table S2).

Fig. 1
figure 1

The violin diagram shows the physical and chemical properties of STP proteins in six species. (A) The number of amino acids of STPs. (B) The molecular weight of STPs. (C) The aromaticity of STPs. (D) The instability index of STPs. (E) The isoelectric point of STPs. (F) The gravy of STPs. The ‘ns’ means no significant difference

Phylogenetic relationships of STP family members from six Brassicaceae species

To analyse the possible evolutionary characteristics of the STP gene family in Brassicaceae, we conducted a phylogenetic tree based on 177 STP amino acid sequences from six species. All STP proteins were clustered into four groups (Fig. 2). In comparison, group II contained the most STP gene family members, followed by group III and group IV, and Group I had the fewest members (Fig. 3). For group III and IV, R. sativus, B. oleracea and B. rapa had the same STP gene number, which was between the STP gene number of A. thaliana and B. napus or B. juncea (Fig. 3). For Group I, R. sativus had only one member because of the loss of STP5.

Fig. 2
figure 2

An NJ phylogenetic tree of STP protein sequences from A. thaliana (At), R. sativus (Rs), B. oleracea (Bo), B. rapa (Br), B. napus (Bn) and B. juncea (Bj). Different colours of branches represent different groups

Fig. 3
figure 3

The STP gene numbers of four groups among different Brassicaceae species, including A. thaliana (At), R. sativus (Rs), B. oleracea (Bo), B. rapa (Br), B. napus (Bn) and B. juncea (Bj)

Conserved motif distribution and structural analysis of the STP family

To gain insight into potential functions and diversification among STPs, the encoded conserved motifs and exon–intron organizations were compared. As expected, most phylogenetically closely related STPs shared similar motifs and structures (Fig. 4). Fifteen predicted motifs were identified throughout the STP protein sequences. Motifs 1, 2, 5, 7, 8, 9, and 14 were present in all analysed STPs. The length of motifs ranged from 15 to 92, and part of the putative sugar_tr domain was predicted in motifs 1–7 and motif 11 (Table 1).

Fig. 4
figure 4

Motif and structural analyses of STPs. (A) Phylogenetic tree of STP proteins. (B) Schematic representation of the conserved motif compositions of STP. (C) Exon/intron structures of STP genes

Table 1 Conserved motifs commonly identified in STP proteins

The exon/intron structures exhibited a highly conserved organization in STP genes. Most STPs (67%) presented four exons divided by three introns, and 28% of STPs had 3 exons. Three percent of STPs had 2 exons, and only 3 STPs had more than 5 exons (Fig. S2).

Tandem duplications and synteny of STP genes

Segmental and tandem duplications provide critical sources of primitive genetic material for genome complexity and evolutionary novelty. We investigated the syntenic and tandem relationships of STP genes. In A. thaliana, AtSTP4 and AtSTP10, distributed on chromosome 1, were demonstrated to be tandem duplications. For the other species, STP4 and STP10 on chromosome 1 were located in tandem duplicated regions. STP10 had 3–5 copies in the tandem duplication cluster. However, this tandem duplication cluster was not found in B. napus.

Of the 14 AtSTPs, most have a syntenic relationship with STP genes in other species, except for AtSTP3, which exhibited no syntenic genes in B. juncea, AtSTP5, which lacked a syntenic gene in R. sativus, AtSTP10, which had no syntenic relationship with B. napus, AtSTP12, which lacked a syntenic gene in R. sativus and B. oleracea (Fig. 5 and Table S1). Additionally, most AtSTP genes were associated with more than one gene pair with other species. For instance, both AtSTP4 and AtSTP6 have more than four syntenic genes in other species. In addition, most STP genes in R. sativus, B. oleracea, B. rapa, B. napus and B. juncea have found syntenic STP genes in A. thaliana (Fig. S3).

Fig. 5
figure 5

Synteny analysis of STP genes between A. thaliana and five Brassicaceae crops. The red lines highlight the syntenic STP gene pairs. (A), (B), (C), (D), and (E) indicate synteny gene pairs between R. sativus, B. oleracea, B. rapa, B. napus, B. juncea and A. thaliana, respectively

Cis-acting element analysis of RsSTPs

To investigate the mechanism of transcriptional control of RsSTP genes, the cis-acting elements in the 1.5 kb potential promoter region of these genes were identified using the PlantCARE program. A total of 49 types of cis-acting elements were identified in the promoter regions of RsSTP genes, except for common cis-acting elements in promoters (e.g., CAAT boxes and TATA boxes) (Fig. S4). The cis-acting elements involved in light responsiveness were most abundant in the promoter regions of RsSTPs, indicating that they might participate in light regulatory pathways. Additionally, cis-acting elements responded to phytohormones such as abscisic acid (ABRE), auxin (TGA-element, AuxRE, TGA-box and AuxRR-core), methyl jasmonate (CGTCA-motif and TGAGG-motif), salicylic acid (TCA-element) and gibberellin (P-box, GARE-box and TATC-box). Furthermore, stress-related cis-acting elements were also found in the promoter regions, including anaerobic (ARE), drought (MBS), low-temperature (LTR), defence and stress (TC-rich repeats) and anoxic (GC motif) regions.

Expression profiles of RsSTPs in different tissues

Gene expression patterns are always associated with functional divergence in a gene family [7, 29]. Therefore, public RNA-seq data were used to analyse the expression patterns of RsSTPs in various tissues (taproot, leaf, bolting, flower, seedpod and callus) [30]. In the present study, the heatmap results revealed that the expression levels of nine radish STP genes (RaSTP1.1/6.3/8/10.1/10.2/10.3/10.4/11/13.1) were very low, with fragments per kilobase of transcript per million mapped reads (FPKM) < 2 in all detected tissues (Fig. 6). RsSTP4.1 and RsSTP4.3 were expressed in all detected tissues, while a large portion of RsSTPs were organ-specific. For example, RsSTP2.1/2.2/4.2/6.1/7.2/9.1/9.2/14 were specifically expressed in flowers. The heatmap also shows the highest expression level for both RsSTP3 and RsSTP6.2 in seedpods and RsSTP13.2 in calli. The diverse expression patterns suggested that RsSTPs might be involved in a variety of biological functions during radish growth and development.

Fig. 6
figure 6

Expression profiles of radish STP genes in various tissues. The fragments per kilobase of transcript per million mapped reads (FPKM) data were log2 transformed. The color bar is shown at the right

Expression profiles of RsSTPs in response to A. tumefaciens infection

To clarify the response of RsSTPs to biotic stress, expression was compared in hypocotyls of susceptible (Lin 19) and resistant (Line 18) inbred lines inoculated with A. tumefaciens for 7 days by public RNA-seq data. The expression of a total of 12 RsSTPs (RsSTP2.1/2.2/4.2/6.1/7.1/7.2/8/9.1/10.1/10.2/10.3/10.4) was undetectable in both Lines 18 and 19 under the treatment and control conditions (Fig. 7). Three genes (RsSTP1.1/1.2/4.1) were not significantly affected, and three genes (RsSTP3/13.1/14) were significantly downregulated by A. tumefaciens infection in both lines. RsSTP4.3 was not significantly affected in Line 18 by A. tumefaciens infection but was significantly downregulated in Line 19, while the expression of RsSTP6.3 and RsSTP7.3 was the opposite. The expression of RsSTP6.2 and RsSTP13.2 was significantly upregulated in Line 18 by A. tumefaciens infection but undetectable in Line 19. These two genes most likely confer Line 18 resistance to A. tumefaciens infection.

Fig. 7
figure 7

The expression profiles of RsSTPs 7 d after incubation with Agrobacterium tumefaciens in radish hypocotyls. Lines 18 and 19 are resistance and susceptible radish inbred lines to A. tumefaciens, respectively. CK and T represent 7 d after incubation with LB medium and A. tumefaciens, respectively. The fragments per kilobase of transcript per million mapped reads (FPKM) data were log2 transformed. The color bar is shown at the right

Expression profiles of RsSTPs in response to cadmium (Cd), chromium (Cr) and lead (Pb) stress

The heatmap results indicated that a large number of RsSTPs showed a very low expression level under the control, and most of them were not activated by heavy metal treatment, except for RsSTP4.1, RsSTP4.3 and RsSTP13.2. For RsSTP4.1 and RsSTP4.3, the expression levels were 2.5- and 15.7-fold higher under Pb stress and 21.3- and 37.4-fold higher under Cd stress than under the control, respectively, but neither was significantly affected by Cr stress. Notably, the expression levels of RsSTP13.2 under Pb, Cd and Cr stress were 64.1-, 148.3- and 412.5-fold higher than the expression levels of RsSTP13.2 under the control, respectively (Fig. 8). Therefore, these three significantly induced genes are most likely candidates confer radish resistance to heavy metal stress.

Fig. 8
figure 8

Expression profiles of RsSTP responses to cadmium (Cd), chromium (Cr) and lead (Pb) stress in radish. The fragments per kilobase of transcript per million mapped reads (FPKM) data were log2 transformed. The color bar is shown at the right

Expression profiles of RsSTPs in response to NaCl and PEG-6000 stress and glucose, sucrose and fructose treatment

A quantitative real-time PCR (RT-qPCR) assay was performed to examine RsSTP expression levels in root tissue under salinity and simulated drought (PEG-6000) stress and their response to exogenous sugar treatment, including glucose, fructose and sucrose (Fig. 9). The RT-qPCR results indicated that RsSTP9.2 and RsSTP14 were not expressed in radish roots of either the treatment or control. The expression of eight genes, including RsSTP2.2, RsSTP3, RsSTP4.1, RsSTP4.3, RsSTP6.1, RsSTP7.1, RsSTP7.3 and RsSTP11, was not significantly affected by any treatment. Three genes showed the same expression pattern in response to NaCl and PEG-6000 treatment, both of which significantly downregulated the expression of RsSTP1.1 but upregulated the expression of RsSTP2.1 and RsSTP13.2. Exogenous sugar treatment significantly affected the expression of RsSTPs in radish roots. All exogenous sugar treatments in the present study significantly repressed the expression of RsSTP1.1, RsSTP9.1 and RsSTP10. In addition, the expression of RsSTP1.2 and RsSTP4.2 was repressed by glucose and fructose, respectively. The expression of RsSTP8 and RsSTP13.1 was significantly upregulated by fructose, and RsSTP1.2 and RsSTP4.2 were significantly downregulated by sucrose.

Fig. 9
figure 9

Quantitative real-time PCR analysis of RsSTP expression levels in the roots in response to 1.5% NaCl and 20% PEG stress and 2% glucose, 2% fructose and 2% sucrose treatment. The presented gene expression levels are relative to the expression of the reference gene RsGAPDH. Data are presented as the mean ± standard error of three independent experiments. CK: control treatment with distilled water

Discussion

STP proteins in plants play vital roles in monosaccharide transport and are involved in regulating multiple growth and developmental processes and biotic and abiotic resistance. In recent years, with the availability of various plant genomes, genome-wide and expression analysis of the STP gene family has been reported in many plants, such as Arabidopsis thaliana [6], Oryza sativa [27, 31], Solanum lycopersicum [32], Pyrus bretschneideri Rehd) [33], Brassica oleracea var. capitata L. [34], Fragaria vesca [35], Capsicum annuum L. [29], Manihot esculenta [36], Triticum aestivum L. [37] and son on. However, the evolutionary dynamics and functional analysis of the STP gene family in Brassicaceae crops are still largely unknown.

In the present study, a total of 25, 25, 28, 36 and 49 STP-encoding genes were identified from the genomes of R. sativus, B. oleracea, B. rapa, B. napus and B. juncea, respectively. The same STP gene numbers were identified in R. sativus and B. oleracea even though R. sativus (460 Mb and 44,109 coding genes, respectively) [38] and B. oleracea (648 Mb and 54,475 coding genes, respectively) [39] have different genome sizes and gene numbers. B. rapa (442.9 Mb and 45,985 coding genes, respectively) [40] and R. sativus have similar genome sizes and gene numbers, but variant in STP gene numbers. Thus, the number of STP genes is no direct correlation with genome size and the total number of coding genes in Brassicaceae crops. The same result was also reported by a previous report in Gramineae crops [7]. Tetraploid B. juncea (920 Mb and 101,959 coding genes, respectively) [41], which have undergone an allopolyploidization event, as well as B. napus (1008 Mb and 100,919 coding genes, respectively) [42], have an almost twice larger genome size and number of coding genes than diploid R. sativus, B. oleracea and B. rapa, and have significantly more STP gene numbers were identified. This result makes us speculated that a large number of STP genes retained after the allopolyploidization event, even though accompanied by gene losses occur in this process.

Segmental and tandem duplication events play a critical role in the expansion and increased functional diversity of the STP gene family. Previous studies indicated that a WGT occurred in the common ancestor of Brassicaceae crops following its divergence from A. thaliana [28]. In this study, we revealed that most AtSTP genes have syntenic pairs in other Brassicaceae species with more than one copy, which is consistent with the polyploidization of these species. Here, segmental duplication was the major force driving the expansion of STP genes in Brassicaceae. Additionally, several genes were lost or retained one copy, suggesting that there may have been some variability in the gene loss events during evolution. Previous studies concluded that functionally redundant genes prefer to be lost in the diploidization process occurring after paleopolyploidy events [43]. The tandem duplication STP10 genes were present in R. sativus, B. rapa, B. oleracea and two subgenomes of B. juncea, indicating that the associated tandem duplication event occurred in the common ancestor.

We revealed physical and chemical properties, diverse gene structures, conserved motifs and phylogenetic analysis of STP genes in Brassicaceae crops. The physical and chemical properties of STPs showed that there were no significant differences among the six species, which indicated that the STP proteins were conserved among different species. Furthermore, most of the STP genes contained four exons and three introns, which is similar to other plant species, such as rice, tomato, and pears [27, 31,32,33]. A phylogenetic analysis revealed that the STPs in Brassicaceae crops were classified into four groups, consistent with the classification in cassava and Gramineae [7, 36], suggesting that STP proteins are highly conserved across lineages. These results indicated that STP functions are mainly conserved in different species.

Previous studies have produced evidence that STPs are involved in plant growth and developmental processes and biotic and abiotic stress responses. The cis-acting elements in the promoter regions are always associated with their transcriptional control and function. In the present study, cis-acting elements involved in light responsiveness, responding to phytohormones and stresses were also found in the predicted promoter regions of RsSTPs (Fig. S4), suggest that these genes might be involved in radish growth and developmental processes and biotic and abiotic stress responses.

In the process of evolution, most of the homologous genes of different species retain the same or similar biological functions, and the expression patterns of genes are often closely related to their functions. In the present study, expression analysis of RsSTPs in different tissues showed that RsSTP4.1 and RsSTP4.3 were expressed in all tissues and their expression levels were significantly induced under Pb and Cd stress, but not by Cr stress and A. tumefaciens infection. AtSTP4, which is orthologous in A. thaliana, could transport a broad spectrum of monosaccharides [11], and induced by fungal biotroph Erysiphe cichoracearum infection [44]. Therefore, RsSTP4.1 and RsSTP4.3 might play vital roles in soluble sugar transporters in different tissues of radish and also involved in Pb and Cd stress resistance. However, their biological function needs further experimental verification. AtSTP4/6/8/9/10/11 were found to be highly expressed in pollen tubes by a previous study [22]. Our present results show that eight genes (RsSTP2.1/2.2/4.2/6.1/7.2/9.1/9.2/14) were specifically expressed in flowers, and we conjectured that these genes might be responsible for glucose uptake into pollen tubes. RsSTP3 and RsSTP6.2 were highly expressed in seedpods, suggesting that they might be involved in soluble sugar accumulation in this tissue. STP genes were reported participate in response to biological stress in other plants. The expression of BoSTP4b and BoSTP12 were up-regulated in cabbage with Plasmodiophora brassicae infection [34]. RsSTP6.2 and RsSTP13.2 most likely confer radish resistance to A. tumefaciens infection by RNA-seq data analysis in the present study. Earlier studies showed that STP genes play important roles in various sugar transportation processes [4, 6, 8,9,10,11,12,13,14,15,16,17,18,19]. We detected that the expression of most genes was not significantly affected by 2% glucose, fructose and sucrose treatment, and five genes were significantly downregulated by at least one treatment (Fig. 9). Only RsSTP8 and RsSTP13.1 were upregulated by fructose treatment. Deng et al. also indicated that only OsST8 was upregulated by fructose treatment in the roots of rice [27]. The transporter of glucose and sucrose in radish roots might be responsible for other sugar transporters.

RsSTP13.2 was significantly upregulated in resistant plants by A. tumefaciens infection but undetectable in susceptible plants and induced by Cd, Cr, Pb, NaCl and PEG-6000 stress, indicating that RsSTP13.2 is involved in resistance to both biotic and abiotic stress in radish. A previous study also indicated that STP13 is involved in biotic and abiotic responses and resistance in other plants. In A. thaliana, STP13 maintains low expression under normal conditions, but STP13 is induced by MYB96 and reabsorbs the monosaccharides that are released by damaged cells under saline conditions [45, 46]. The upregulation of AtSTP13 deprivation of extracellular sugar levels, which is used as an energy source for pathogens, enhances antibacterial defence [47]. In addition, AtSTP13 is also involved in resistance to Botrytis cinerea by affecting glucose transport [23]. In Phaseolus vulgaris L., both excess zinc and iron/zinc deficient stress significantly upregulate the expression of STP13 in the shoot, while excess zinc stress induced and iron/zinc deficient stress decreased its expression in roots [26].

Conclusions

The present study provides insights into the evolutionary patterns of the STP gene family in Brassicaceae genomes and provides a theoretical basis for future functional analysis of STP genes in Brassicaceae crops. RsSTP13.2 may serve as a candidate gene to improve the biotic and abiotic resistance of plants through transgenic technology.

Methods

Data sources and identification of RsSTP family genes

The R. sativus ‘Xin-li-mei’ whole-genome sequence [38] was used to identify the RsSTP family genes. The whole-genome sequences of B. napus [42], B. juncea [41], B. rapa [40], B. oleracea [39] and A. thaliana were obtained from the BNPIR database (http://cbi.hzau.edu.cn/cgi-bin/rape/download_ext), NCBI database (https://www.ncbi.nlm.nih.gov/bioproject/PRJNA550308), BRAD database (http://brassicadb.cn/#/), Ensembl (http://plants.ensembl.org/info/data/ftp/index.html) and TAIR database (http://www.arabidopsis.org/), respectively. Hmmsearch integrated in HMMER software [48] was used to identify STP gene family members with the hidden Markov model (HMM) profile of Sugar_tr (PF00083) obtained from the Pfam database (http://pfam.xfam.org). Then, the SMART database (http://smart.embl-heidelberg.de) and Batch CD-Search (https://www.ncbi.nlm.nih.gov/Structure/bwrpsb/bwrpsb.cgi) were further used to confirm the genes that were obtained with highly conserved sugar_tr domains and MFS_STP domains. The STP genes of B. napus, B. juncea, B. rapa, B. oleracea and R. sativus were named according to their phylogenetic relationship with A. thaliana STP genes.

The phylogenetic tree construction

For phylogenetic tree construction, first, all the STP full-length protein sequences of six Brassicaceae species were aligned by using the MUSCLE program [49]. Then, MEGAX [50] was used to construct a neighbour-joining tree using the Jones-Taylor-Tornton (JTT) model with 1000 bootstrap replicates. Additionally, uniform rates and homogeneous lineages were adopted, and partial deletion with a site coverage threshold of 70% was given for gaps/missing data.

Sequence properties, conserved motifs and gene structure analyses

The Biopython module Bio.SeqUtils. ProtParam of Python language was used to calculate the molecular weight, aromaticity and other physical and chemical properties of STP proteins. The R script was used to compare the differences among different species using ggpubr with ‘anova’ methods. The MEME suite [51] was used to identify the conserved motifs of STP proteins with the following parameters: the maximum number to be found was set to 15, and the motif length was set to 8–100 bp. The Pfam domains of motifs were identified in the Pfam database (http://pfam.xfam.org). The gene structures information containing extron and intron position were obtained from GFF file using python script. The TBtools program [52] was used to visualize gene structures and conserved motifs.

Tandem duplications and syntenic analysis of STP genes

Tandem genes in A. thaliana and other species were defined as those genes that were separated by ten or fewer genes. The Multiple Collinearity Scan toolkit (MCScanX) [53] was used to identify syntenic duplication events between A. thaliana and other species, with the default parameters. The synteny relationship of STP genes was visualized using TBtools software [52]. The subgenomes of B. napus and B. juncea were calculated separately.

Cis-acting element analyses of RsSTPs and its transcriptional profiles in RNA-seq data

The cis-acting elements in the potential promoter region (the upstream 1.5 kb sequence starting from the start codon) of the radish STP genes were identified using the PlantCARE program (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/).

The RsSTP expression profiles of different tissues (root, leaf, bolting, flower, silique and callus), their response to heavy metals (cadmium, chromium and lead) and Agrobacterium tumefaciens infection were analysed based on public transcriptome data [30, 54,55,56,57]. The fragments per kilobase of transcript per million mapped reads (FPKM) data were log2 transformed, and a heatmap was created using TBtools [52].

Plant materials and stress treatments

Seeds of radish cultivar ‘Xin-li-mei’ purchased from Jingyan Yinong (Beijing) Seed Sci-Tech Co., Ltd. were surface-sterilized in 1% NaClO and incubated at 22 °C for 2 d in darkness. The germinated seeds were sown into plastic pots and incubated in a growth chamber at a 16 h day (22 °C)/8 h night (20 °C) cycle. Seedlings were watered as needed with half-strength Hoagland’s nutrient solution. Plants at the three true leaf stages were used for subsequent abiotic stress and sugar treatment. For the simulated salinity and drought stress treatments, the seedlings were subjected to 1.5% NaCl and 20% PEG-6000 for 3 h, respectively. For sugar treatments, the seedlings were subjected to 2.0% glucose, sucrose and fructose solution for 3 h. The seedlings treated with sterile water were used as a control. Eight plants were used as biological replicates, and the roots were collected. The samples were frozen in liquid nitrogen immediately after collection and stored at -80 °C for RNA extraction. The collection of plant material, is in compliance with relevant institutional, national, and international guidelines and legislation.

RNA extraction, reverse transcription quantitative polymerase chain reaction (RT-qPCR) analysis of RsSTPs

Total RNA was extracted from different radish samples with a Quick RNA Isolation Kit (Huayueyang, Beijing, China) according to the manufacturer's instructions. A total of 800 ng of high-quality total RNA was used to synthesize first-strand cDNA with PrimeScript® Reverse Transcriptase (Takara Biotechnology, Dalian, China). The SYBR Green qPCR kit (Takara Biotechnology) was used for RT-qPCR in a Stratagene Mx3000P thermocycler (Agilent, Santa Clara, CA, USA). RsGAPDH was utilized as an internal control, and the relative expression levels of RsSTPs were calculated with the 2−ΔΔCt method [58, 59]. The primer used for RT-qPCR is shown in Table S1.

Availability of data and materials

The R. sativus ‘Xin-li-mei’ whole-genome sequence was downloaded from CNCB database with accession number of GWHANWD00000000 (https://ngdc.cncb.ac.cn/gwh/Assembly/9797/show). The genome and its annotation file of B. napus version ZS11 were obtained from BnPIR database (http://cbi.hzau.edu.cn/cgi-bin/rape/download_ext). The genome and its annotation file of B. juncea were obtained from NCBI with the accession PRJNA550308 (https://www.ncbi.nlm.nih.gov/bioproject/PRJNA550308). The B. rapa genome version 3.5 information was downloaded from BRAD (http://brassicadb.cn/#/Download/). The B. oleracea genome was downloaded from Ensembl (http://plants.ensembl.org/Brassica_oleracea/Info/Index). The A. thaliana genome was obtained from TAIR database (http://www.arabidopsis.org/). The RNA-seq reads in different tissues (root, leaf, bolting, flower, silique and callus) of R. sativus are available at NCBI Sequence Read Archive (SRA, http://www.ncbi.nlm.nih.gov/Traces/sra) (PRJNA413464). The RNA-seq raw data of control and heavy metal stress treatments were downloaded from SRA with accession numbers: SRX256970 (CK), SRX862647 (Cr600), SRX263753 (Pb1000) and SRX824523 (Cd200). The RNA-seq raw data of control and Agrobacterium tumefaciens treatments were downloaded from SRA with accession numbers: SRX9205566 (control 18 line), SRX9205567 (tumor 18 line), SRX9205568 (control 19 line) and SRX9205569 (tumor 19 line).

Abbreviations

STPs:

Sugar Transporter Proteins

SWEETs:

Sugars Will Eventually Be Exported Transporters

MSTs:

Monosaccharide Transporters

SUTs:

Sucrose Transporters

VGT:

Vacuolar Glucose Transporter

TMT:

Tonoplast Monosaccharide Transporter

GlcT:

Plastidic Glucose Transporter

SGB1:

Suppressor of G Protein Beta 1

PLT:

Polyol Transporter

INT:

Inositol Transporter

ERD6L:

Early Response to Dehydration-6-Like

Cd:

Cadmium

Cr:

Chromium

Pb:

Lead

JTT:

Jones-Taylor-Tornton

MCScanX:

Multiple Collinearity Scan toolkit

RT-qPCR:

Reverse Transcription Quantitative Polymerase Chain Reaction

References

  1. Geng Y, Dong X, Zhang C. Recent Progress of Sugar Transporters in Horticultural Crops. Acta Horticulturae Sinica. 2021;48(4):676–88.

    Google Scholar 

  2. Hellmann H, Smeekens S. Sugar sensing and signaling in plants. Front Plant Sci. 2014;5:113.

    PubMed  PubMed Central  Google Scholar 

  3. Lastdrager J, Hanson J, Smeekens S. Sugar signals and the control of plant growth and development. J Exp Bot. 2014;65(3):799–807.

    Article  CAS  PubMed  Google Scholar 

  4. Buttner M. The Arabidopsis sugar transporter (AtSTP) family: an update. Plant Biol (Stuttg). 2010;12(Suppl 1):35–41.

    Article  CAS  Google Scholar 

  5. Doidy J, Grace E, Kühn C, Simon-Plas F, Casieri L, Wipf D. Sugar transporters in plants and in their interactions with fungi. Trends Plant Sci. 2012;17(7):413–22.

    Article  CAS  PubMed  Google Scholar 

  6. Buttner M. The monosaccharide transporter(-like) gene family in Arabidopsis. FEBS Lett. 2007;581(12):2318–24.

    Article  PubMed  CAS  Google Scholar 

  7. Kong W, An B, Zhang Y, Yang J, Li S, Sun T, Li Y. Sugar Transporter Proteins (STPs) in Gramineae Crops: Comparative Analysis, Phylogeny, Evolution, and Expression Profiling. Cells. 2019;8(6):560.

    Article  CAS  PubMed Central  Google Scholar 

  8. Büttner M, Truernit E, Baier K, Scholz-Starke J, Sontheim M, Lauterbach C, Huss VAR, Sauer N. AtSTP3, a green leaf-specific, low affinity monosaccharide-H+ symporter of Arabidopsis thaliana. Plant, Cell Environ. 2000;23(2):175–84.

    Article  Google Scholar 

  9. Sauer N, Friedländer K, Gräml-Wicke U. Primary structure, genomic organization and heterologous expression of a glucose transporter from Arabidopsis thaliana. EMBO J. 1990;9(10):3045–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Truernit E, Stadler R, Baier K, Sauer N. A male gametophyte-specific monosaccharide transporter in Arabidopsis. Plant J. 1999;17(2):191–201.

    Article  CAS  PubMed  Google Scholar 

  11. Truernit E, Schmid J, Epple P, Illig J, Sauer N. The sink-specific and stress-regulated Arabidopsis STP4 gene: enhanced expression of a gene encoding a monosaccharide transporter by wounding, elicitors, and pathogen challenge. Plant Cell. 1996;8(12):2169–82.

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Scholz-Starke J, Büttner M, Sauer N. AtSTP6, a new pollen-specific H+-monosaccharide symporter from Arabidopsis. Plant Physiol. 2003;131(1):70–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Schneidereit A, Scholz-Starke J, Sauer N, Büttner M. AtSTP11, a pollen tube-specific monosaccharide transporter in Arabidopsis. Planta. 2005;221(1):48–55.

    Article  CAS  PubMed  Google Scholar 

  14. Norholm MH, Nour-Eldin HH, Brodersen P, Mundy J, Halkier BA. Expression of the Arabidopsis high-affinity hexose transporter STP13 correlates with programmed cell death. FEBS Lett. 2006;580(9):2381–7.

    Article  CAS  PubMed  Google Scholar 

  15. Schneidereit A, Scholz-Starke J, Büttner M. Functional characterization and expression analyses of the glucose-specific AtSTP9 monosaccharide transporter in pollen of Arabidopsis. Plant Physiol. 2003;133(1):182–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Rottmann T, Klebl F, Schneider S, Kischka D, Rüscher D, Sauer N, Stadler R. Sugar transporter STP7 specificity for l-Arabinose and d-Xylose contrasts with the typical hexose transporters STP8 and STP12. Plant Physiol. 2018;176(3):2330–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Liu J, Liu M, Tan L, Huai B, Ma X, Pan Q, Zheng P, Wen Y, Zhang Q, Zhao Q, et al. AtSTP8, an endoplasmic reticulum-localised monosaccharide transporter from Arabidopsis, is recruited to the extrahaustorial membrane during powdery mildew infection. New Phytol. 2021;230(6):2404–19.

    Article  CAS  PubMed  Google Scholar 

  18. Rottmann T, Zierer W, Subert C, Sauer N, Stadler R. STP10 encodes a high-affinity monosaccharide transporter and is induced under low-glucose conditions in pollen tubes of Arabidopsis. J Exp Bot. 2016;67(8):2387–99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Poschet G, Hannich B, Büttner M. Identification and characterization of AtSTP14, a novel galactose transporter from Arabidopsis. Plant Cell Physiol. 2010;51(9):1571–80.

    Article  CAS  PubMed  Google Scholar 

  20. Otori K, Tanabe N, Tamoi M, Shigeoka S. Sugar Transporter Protein 1 (STP1) contributes to regulation of the genes involved in shoot branching via carbon partitioning in Arabidopsis. Biosci Biotechnol Biochem. 2019;83(3):472–81.

    Article  CAS  PubMed  Google Scholar 

  21. Flütsch S, Nigro A, Conci F, Fajkus J, Thalmann M, Trtílek M, Panzarová K, Santelia D. Glucose uptake to guard cells via STP transporters provides carbon sources for stomatal opening and plant growth. EMBO Rep. 2020;21(8): e49719.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Rottmann T, Fritz C, Sauer N, Stadler R. Glucose uptake via STP transporters inhibits in vitro pollen tube growth in a HEXOKINASE1-dependent manner in Arabidopsis thaliana. Plant Cell. 2018;30(9):2057–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Lemonnier P, Gaillard C, Veillet F, Verbeke J, Lemoine R, Coutos-Thévenot P, La Camera S. Expression of Arabidopsis sugar transport protein STP13 differentially affects glucose transport activity and basal resistance to Botrytis cinerea. Plant Mol Biol. 2014;85(4–5):473–84.

    Article  CAS  PubMed  Google Scholar 

  24. Huai B, Yang Q, Wei X, Pan Q, Kang Z, Liu J. TaSTP13 contributes to wheat susceptibility to stripe rust possibly by increasing cytoplasmic hexose concentration. BMC Plant Biol. 2020;20: 49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Huai B, Yang Q, Qian Y, Qian W, Kang Z, Liu J. ABA-Induced Sugar Transporter TaSTP6 Promotes Wheat Susceptibility to Stripe Rust. Plant Physiol. 2019;181(3):1328–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Urwat U, Zargar SM, Ahmad SM, Ganai NA. Insights into role of STP13 in sugar driven signaling that leads to decrease in photosynthesis in dicot legume crop model (Phaseolus vulgaris L.) under Fe and Zn stress. Mol Biol Rep. 2021;48(3):2527–31.

    Article  CAS  PubMed  Google Scholar 

  27. Deng X, An B, Zhong H, Yang J, Kong W, Li Y. A Novel Insight into Functional Divergence of the MST Gene Family in Rice Based on Comprehensive Expression Patterns. Genes. 2019;10(3):239.

    Article  CAS  PubMed Central  Google Scholar 

  28. Cheng F, Liang J, Cai C, Cai X, Wu J, Wang X. Genome sequencing supports a multi-vertex model for Brassiceae species. Curr Opin Plant Biol. 2017;36:79–87.

    Article  CAS  PubMed  Google Scholar 

  29. Wei H, Liu J, Zheng J, Zhou R, Cheng Y, Ruan M, Ye Q, Wang R, Yao Z, Zhou G, et al. Sugar transporter proteins in Capsicum: identification, characterization, evolution and expression patterns. Biotechnol Biotechnol Equip. 2020;34(1):341–53.

    Article  CAS  Google Scholar 

  30. Wang J, Qiu Y, Wang X, Yue Z, Yang X, Chen X, Zhang X, Shen D, Wang H, Song J, et al. Insights into the species-specific metabolic engineering of glucosinolates in radish (Raphanus sativus L.) based on comparative genomic analysis. Scientific Rep. 2017;7(1):16040.

    Article  CAS  Google Scholar 

  31. Toyofuku K, Kasahara M, Yamaguchi J. Characterization and expression of monosaccharide transporters (osMSTs) in rice. Plant Cell Physiol. 2000;41(8):940–7.

    Article  CAS  PubMed  Google Scholar 

  32. Reuscher S, Akiyama M, Yasuda T, Makino H, Aoki K, Shibata D, Shiratake K. The sugar transporter inventory of tomato: genome-wide identification and expression analysis. Plant Cell Physiol. 2014;55(6):1123–41.

    Article  CAS  PubMed  Google Scholar 

  33. Li JM, Zheng DM, Li LT, Qiao X, Wei SW, Bai B, Zhang SL, Wu J. Genome-Wide Function, Evolutionary Characterization and Expression Analysis of Sugar Transporter Family Genes in Pear (Pyrus bretschneideri Rehd). Plant Cell Physiol. 2015;56(9):1721–37.

    Article  CAS  PubMed  Google Scholar 

  34. Zhang W, Wang S, Yu F, Tang J, Yu L, Wang H, Li J. Genome-Wide Identification and Expression Profiling of Sugar Transporter Protein (STP) Family Genes in Cabbage (Brassica oleracea var. capitata L.) Reveals their Involvement in Clubroot Disease Responses. Genes. 2019;10(1):71.

    Article  PubMed Central  CAS  Google Scholar 

  35. Jiu S, Haider MS, Kurjogi MM, Zhang K, Zhu X, Fang J. Genome-wide Characterization and Expression Analysis of Sugar Transporter Family Genes in Woodland Strawberry. Plant Genome. 2018;11(3):170103.

  36. Liu Q, Dang H, Chen Z, Wu J, Chen Y, Chen S, Luo L. Genome-Wide Identification, Expression, and Functional Analysis of the Sugar Transporter Gene Family in Cassava (Manihot esculenta). Int J Mol Sci. 2018;19(4):987.

    Article  PubMed Central  CAS  Google Scholar 

  37. Liu H, Li C, Qiao L, Hu L, Wang X, Wang J, Ruan X, Yang G, Yin G, Wang C, et al. The Sugar Transporter family in wheat (Triticum aestivum L.): genome-wide identification, classification, and expression profiling during stress in seedlings. PeerJ. 2021;9:e11371.

    Article  PubMed  PubMed Central  Google Scholar 

  38. Zhang X, Liu T, Wang J, Wang P, Qiu Y, Zhao W, Pang S, Li X, Wang H, Song J, et al. Pan-genome of Raphanus highlights genetic variation and introgression among domesticated, wild, and weedy radishes. Mol Plant. 2021;14(12):2032–55.

    Article  CAS  PubMed  Google Scholar 

  39. Parkin IA, Koh C, Tang H, Robinson SJ, Kagale S, Clarke WE, Town CD, Nixon J, Krishnakumar V, Bidwell SL, et al. Transcriptome and methylome profiling reveals relics of genome dominance in the mesopolyploid Brassica oleracea. Genome Biol. 2014;15(6):R77.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Zhang L, Cai X, Wu J, Liu M, Grob S, Cheng F, Liang J, Cai C, Liu Z, Liu B, et al. Improved Brassica rapa reference genome by single-molecule sequencing and chromosome conformation capture technologies. Horticulture research. 2018;5:50–50.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. Paritosh K, Yadava SK, Singh P, Bhayana L, Mukhopadhyay A, Gupta V, Bisht NC, Zhang J, Kudrna DA, Copetti D, et al. A chromosome-scale assembly of allotetraploid Brassica juncea (AABB) elucidates comparative architecture of the A and B genomes. Plant Biotechnol J. 2021;19(3):602–14.

    Article  CAS  PubMed  Google Scholar 

  42. Song J-M, Guan Z, Hu J, Guo C, Yang Z, Wang S, Liu D, Wang B, Lu S, Zhou R, et al. Eight high-quality genomes reveal pan-genome architecture and ecotype differentiation of Brassica napus. Nature Plants. 2020;6(1):34–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Tang H, Bowers JE, Wang X, Ming R, Alam M, Paterson AH. Synteny and Collinearity in Plant Genomes. Science (New York, NY). 2008;320(5875):486–8.

    Article  CAS  Google Scholar 

  44. Fotopoulos V, Gilbert MJ, Pittman JK, Marvier AC, Buchanan AJ, Sauer N, Hall JL, Williams LE. The monosaccharide transporter gene, AtSTP4, and the cell-wall invertase, Atβfruct1, are induced in Arabidopsis during infection with the fungal biotroph Erysiphe cichoracearum. Plant Physiol. 2003;132(2):821–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Yamada K, Kanai M, Osakabe Y, Ohiraki H, Shinozaki K, Yamaguchi-Shinozaki K. Monosaccharide Absorption Activity of Arabidopsis Roots Depends on Expression Profiles of Transporter Genes under High Salinity Conditions. J Biol Chem. 2011;286(50):43577–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Lee HG, Seo PJ. Transcriptional activation of SUGAR TRANSPORT PROTEIN 13 mediates biotic and abiotic stress signaling. Plant Signal Behav. 2021;16(8):1920759.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. Yamada K, Saijo Y, Nakagami H, Takano Y. Regulation of sugar transporter activity for antibacterial defense in Arabidopsis. Science (New York, NY). 2016;354(6318):1427–30.

    Article  CAS  Google Scholar 

  48. Finn RD, Clements J, Eddy SR. HMMER web server: interactive sequence similarity searching. Nucleic Acids Res. 2011;39(Web Server issue):W29-37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004;32(5):1792–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: Molecular Evolutionary Genetics Analysis across Computing Platforms. Mol Biol Evol. 2018;35(6):1547–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Bailey TL, Boden M, Buske FA, Frith M, Grant CE, Clementi L, Ren J, Li WW, Noble WS. MEME SUITE: tools for motif discovery and searching. Nucleic Acids Res. 2009;37(Web Server issue):W202-208.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Chen C, Chen H, Zhang Y, Thomas HR, Xia R. TBtools: An Integrative Toolkit Developed for Interactive Analyses of Big Biological Data. Mol Plant. 2020;13(8):1194–202.

    Article  CAS  PubMed  Google Scholar 

  53. Wang Y, Tang H, DeBarry JD, Tan X, Li J, Wang X. Lee T-h, Jin H, Marler B, Guo H: MCScanX: a toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Res. 2012;40(7):e49–e49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Tkachenko AA, Gancheva MS, Tvorogova VE, Danilov LG, Predeus AV, Dodueva IE, Lutova LL. Transcriptome analysis of crown gall in radish (Raphanus sativus L.) inbred lines. Annals Applied Biol. 2021;178(3):527–48.

    Article  CAS  Google Scholar 

  55. Xu L, Wang Y, Liu W, Wang J, Zhu X, Zhang K, Yu R, Wang R, Xie Y, Zhang W, et al. De novo sequencing of root transcriptome reveals complex cadmium-responsive regulatory networks in radish (Raphanus sativus L.). Plant Sci. 2015;236:313–23.

    Article  CAS  PubMed  Google Scholar 

  56. Xie Y, Ye S, Wang Y, Xu L, Zhu X, Yang J, Feng H, Yu R, Karanja B, Gong Y, et al. Transcriptome-based gene profiling provides novel insights into the characteristics of radish root response to Cr stress with next-generation sequencing. Front Plant Sci. 2015;6:202.

    PubMed  PubMed Central  Google Scholar 

  57. Wang Y, Xu L, Chen Y, Shen H, Gong Y, Limera C, Liu L. Transcriptome profiling of radish (Raphanus sativus L.) root and identification of genes involved in response to lead (Pb) stress with next generation sequencing. PLoS One. 2013;8(6):e66539.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2−ΔΔCT Method. Methods (San Diego, Calif). 2001;25(4):402–8.

    Article  CAS  Google Scholar 

  59. Liu T, Zhang Y, Zhang X, Sun Y, Wang H, Song J, Li X. Transcriptome analyses reveal key genes involved in skin color changes of 'Xinlimei' radish taproot. Plant Physiol Biochem. 2019;139:528–39.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by the National Natural Science Foundation of China (grant number 31801858); China Agriculture Research System (grant number CARS-23-A09); New Variety Breeding Project of the Major Science Technology Projects of Zhejiang (grant number 2021C02065); Research Foundation for Talented Scholars of Jinling Institute of Technology (grant number jit-b-202009); Research Foundation Incubation Project of Jinling Institute of Technology (grant number jit-fhxm-202113).

Author information

Authors and Affiliations

Authors

Contributions

T. L and J. W designed and conceived the experiments; Q. B and C. W performed the experiments; T. L and T. H participated in plant material growth and experiments; C. B, J. W and T. L analyzed the data; T. L and J. W wrote the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Jinglei Wang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no conflicts of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1: Figure S1.

Distribution of STP genes on chromosomes. The line on the green bars indicates the location of STP genes on chromosomes. The left values corresponding to the scales indicate physical chromosomes. The red genes indicated tandem duplicated genes.

Additional file 2: Figure S2.

The percent of exons numbers of STP genes.

Additional file 3: Figure S3.

The percentage of segmental or tandem duplication genes in five Brassicaceae crops.

Additional file 4: Figure S4.

The types and numbers of cis-acting elements identified in the promoters of RsSTP genes.

Additional file 5: Table S1.

The identified STP genes in six Brassicaceae species. Table S2. The physical and chemical properties of STP proteins.

Additional file 6: Table S3.

Primers used for RT-qPCR in the present study.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, T., Bao, C., Ban, Q. et al. Genome-wide identification of sugar transporter gene family in Brassicaceae crops and an expression analysis in the radish. BMC Plant Biol 22, 245 (2022). https://doi.org/10.1186/s12870-022-03629-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12870-022-03629-2

Keywords