Skip to main content
  • Research article
  • Open access
  • Published:

Genome-wide systematic characterization of the HAK/KUP/KT gene family and its expression profile during plant growth and in response to low-K+ stress in Saccharum

Abstract

Background

Plant genomes contain a large number of HAK/KUP/KT transporters, which play important roles in potassium uptake and translocation, osmotic potential regulation, salt tolerance, root morphogenesis and plant development. Potassium deficiency in the soil of a sugarcane planting area is serious. However, the HAK/KUP/KT gene family remains to be characterized in sugarcane (Saccharum).

Results

In this study, 30 HAK/KUP/KT genes were identified in Saccharum spontaneum. Phylogenetics, duplication events, gene structures and expression patterns were analyzed. Phylogenetic analysis of the HAK/KUP/KT genes from 15 representative plants showed that this gene family is divided into four groups (clades I-IV). Both ancient whole-genome duplication (WGD) and recent gene duplication contributed to the expansion of the HAK/KUP/KT gene family. Nonsynonymous to synonymous substitution ratio (Ka/Ks) analysis showed that purifying selection was the main force driving the evolution of HAK/KUP/KT genes. The divergence time of the HAK/KUP/KT gene family was estimated to range from 134.8 to 233.7 Mya based on Ks analysis, suggesting that it is an ancient gene family in plants. Gene structure analysis showed that the HAK/KUP/KT genes were accompanied by intron gain/loss in the process of evolution. RNA-seq data analysis demonstrated that the HAK/KUP/KT genes from clades II and III were mainly constitutively expressed in various tissues, while most genes from clades I and IV had no or very low expression in the tested tissues at different developmental stages. The expression of SsHAK1 and SsHAK21 was upregulated in response to low-K+ stress. Yeast functional complementation analysis revealed that SsHAK1 and SsHAK21 could rescue K+ uptake in a yeast mutant.

Conclusions

This study provided insights into the evolutionary history of HAK/KUP/KT genes. HAK7/9/18 were mainly expressed in the upper photosynthetic zone and mature zone of the stem. HAK7/9/18/25 were regulated by sunlight. SsHAK1 and SsHAK21 played important roles in mediating potassium acquisition under limited K+ supply. Our results provide valuable information and key candidate genes for further studies on the function of HAK/KUP/KT genes in Saccharum.

Background

Potassium is an essential mineral nutrient for plant growth and development and is also the most abundant monovalent cation in plants, accounting for approximately 2~10% of plant dry weight [1]. Potassium is involved in many important physiological and biochemical processes, such as cell turgor regulation, cell charge balance regulation, enzyme activity regulation and protein synthesis [1]. Symptoms of plant potassium deficiency usually manifest as weak stems, easy lodging, decreased tolerance to drought and cold and yellow leaves, due to the degradation of proteins and chlorophyll, which leads to tissue necrosis [2]. Thus, potassium is of great importance for improving crop yield and quality. Sugarcane is an important sugar and energy crop with a long growth period, large biomass and large amount of potassium fertilizer absorption. On the one hand, it is estimated that sugarcane needs to absorb approximately 2~2.5 kg of potassium to produce one ton of sugar [3, 4]. On the other hand, sugarcane is mainly cultivated in subtropical and tropical regions, where soil acidification and potassium leaching are common. The contents of total potassium and available potassium in the cultivated layer of these sugarcane areas are low.

Plant cells maintain a relatively high and stable K+ concentration (approximately 100~150 mM) in the cytosol, while the K+ concentration is highly variable in the range of 0.01~1 mM [5]. It is generally believed that there are two mechanisms for potassium uptake by plants, namely, a high-affinity transport system (HATS) via potassium transporters at low external potassium concentrations (< 0.2 mM) and a low-affinity transport system (LATS) via potassium channels at high potassium concentrations (> 0.5 mM) [6, 7]. According to their structure and function, potassium transporters in plants can be divided into five families: (1) Shaker channels; (2) TPK (tandem-pore K+) channels; (3) HAK (high-affinity K+ transporter)/KUP (K+ uptake permease)/KT (K+ transporter); (4) HKT transporters; and (5) CPAs (cation-proton antiporters) [2, 8]. Among them, the HAK/KUP/KT family is the largest and is widely distributed in bacteria, fungi and plants but has not been identified in animal cells [9].

According to their homology with bacterial KUP and fungal HAK transporters [10], the plant HAK/KUP/KT transporter members AtKUP1 and HvHAK1 were first cloned from Arabidopsis and barley [11, 12]. Both genes could complement K+ uptake-deficient strains of yeast, indicating that they had potassium transporter activity. Subsequently, several HAK/KUP/KT members were cloned and identified, such as AtKUP3 and AtHAK5 in Arabidopsis, OsHAK1 in rice and CaHAK1 in pepper, which were also shown to be highly compatible potassium transporters [13,14,15,16]. Based on comparative genomic methods, 13, 27 and 27 HAK/KUP/KT genes were identified in Arabidopsis, rice and maize, respectively [17,18,19]. These predicted HAK/KUP/KT transporters were sorted into four clusters. HAK/KUP/KT K+ transporters play versatile roles in potassium ion acquisition and transport, salt stress, osmotic regulation, and the morphogenesis of root and phenotype in plants [7]. The expression of OsHAK1 was greatly induced in the roots of K+-starved rice, while OsHAK5 was less expressed in roots but abundantly expressed in shoots [20, 21]. Some ions, particularly Na+ and NH4+, can have additional effects on the expression of HAK/KUP/KT genes [22, 23].

The transcriptional regulation of HAK/KUP/KT K+ transporters is a universal mechanism by which different plant species respond to K+-starvation stress [8]. The HAK/KUP/KT genes in clade I, such as AtHAK5, OsHAK1, CaHAK1 and ThHAK5, display low expression levels both in roots and shoots under control conditions and are highly upregulated in roots upon K+-deficiency stress [12,13,14, 16]. While the HAK/KUP/KT K+ transporters in other three clades exhibit different expression patterns [24], since transcription of most K+ transporters are not induced by K+ starvation [25]. In Arabidopsis, several transcription factors, including bHLH121 (basic helix-loop-helix 121), DDF2 (dwarf and delayed flowering 2), JLO (jagged lateral organs) and TFII_A (transcription initiation factor II_A gamma chain), have been identified to bind the promoters of HAK5 and activate its expression under low K+ stress [26]. Activation of HAK/KUP/KT K+ transporters is also regulated at posttranscriptional and/or posttranslational level. AtHAK5 and its homologs from pepper and tomato can be activated by the CIPK23 (CBL-interacting protein kinase 23)/CBL (calcineurin B-like protein) complex [27].

In summary, numerous studies have been performed in the functional research of plant HAK/KUP/KT potassium transporters, and important progress has been made. However, the known functional HAK/KUP/KT genes have mainly been identified in a few plants, such as Arabidopsis, rice and maize, but their physiological functions and regulatory mechanisms in sugarcane remain unknown. In this study, based on the newly released S. spontaneum genome [28], we identified the HAK/KUP/KT gene family in S. spontaneum. Phylogenetic relationships among different species, exon/intron organization and gene expression were analyzed. Altogether, these results provide valuable information and robust candidate genes for future functional analyses for the genetic improvement of potassium-utilization efficiency in sugarcane.

Results

Identification of HAK genes in sugarcane

Based on comparative genomics, 29 SbHAK genes were identified from sorghum (Sorghum bicolor, sugarcane’s nearest relative). Using the protein sequences of sorghum HAK genes as a reference, 30 distinct S. spontaneum HAK genes (Table 1), excluding alleles, were identified from the genome of tetraploid S. spontaneum AP85–441 [28]. Each of these genes contained one to four alleles, with an average of 3 (Additional file 1). The 30 SsHAK genes were distributed on seven S. spontaneum chromosomes: chromosome 1 contained six genes; chromosome 2 contained seven genes; chromosome 3 contained four genes; chromosome 4 contained two genes; chromosome 5 contained five genes; and chromosome 6 and 8 each contained three genes. No SsHAK genes were identified on chromosome 7 (Additional file 1).

Table 1 Overview and comparison of HAK genes in Saccharum spontaneum and Sorghum bicolor

All 30 predicted SsHAK proteins had a typical “K_trans” domain (PF02705), which is specific to HAK/KUP/KT potassium transporter family members. For consistency, these SsHAK genes were named based on the previously reported O. sativa HAK nomenclature and phylogenetic relationships [17]. If two SsHAK genes were equally close to a single OsHAK gene, then the same name was used, followed by the letters “a” and “b” (Table 1). Two paralogous SsHAK genes (SsHAK19a and SsHAK19b) were identified that corresponded to the same sorghum gene, Sobic.006G062100, which may be caused by gene loss in sorghum or gene duplication in sugarcane. The number of amino acids in the 30 identified SsHAKs ranged from 487 to 967, with an average of 758. The predicted isoelectric points (pI) of the SsHAKs varied from 5.88 to 9.26, and the average pI was 8.15. The molecular weight ranged from 55.84 kDa to 106.49 kDa, with an average of 84.47 kDa (Table 1). The prediction of transmembrane domains in the SsHAK proteins indicated that most contained 11 or 12 transmembrane helices, which was similar to the findings in sorghum. The subcellular locations of the SsHAK proteins predicted by WoLF PSORT were mainly the plasma membrane, which is most suitable for their roles as transporters to maintain K+ homeostasis in sugarcane. In addition, the SsHAK proteins were also located on some organelles, including the endoplasmic reticulum, vacuole, cytoplasm, Golgi body and chloroplast. Protein sequence alignment of SsHAKs with their orthologs in sorghum showed that S. spontaneum and Sorghum bicolor shared identities ranging from 81 to 98%, with an average of 92.5% (Table 1). Four hundred thirty-five pairwise protein sequence comparisons among these SsHAKs showed that SsHAK19a and SsHAK19b shared the highest identity (96%), while other gene pairs had protein sequence similarities ranging from 28 to 82% with an average of 46%, indicating that the SsHAKs are an ancient gene family with high sequence divergence (Additional file 2).

The nonsynonymous to synonymous substitution ratios (Ka/Ks) between SsHAKs and their orthologous genes in sorghum were calculated to study the evolutionary functional constraints in sugarcane. The results showed that the Ka/Ks ratios were less than 0.5, except for SsHAK13, suggesting that purifying selection was the main force driving the evolution of HAK genes (Fig. 1).

Fig. 1
figure 1

Nonsynonymous (Ka) and synonymous (Ks) substitution ratios of SsHAKs and their orthologs in sorghum. The Ka/Ks ratio was calculated by the Easy_KaKs calculation program (https://github.com/tangerzhang/FAFUcgb/tree/master/easy_Ka-Ks)

Phylogenic analysis of HAK genes in S. spontaneum and representative angiosperms

To analyze the evolution of the HAK gene family in S. spontaneum and different plants, a total of 278 HAK genes from 14 representative angiosperms and a HAK member from Chlamydomonas reinhardtii as the outgroup were used to construct a phylogenetic tree using the neighbor-joining method (Fig. 2, Additional file 3). The 278 HAK genes included 6 from Amborella trichopoda, 8 from Solanum lycopersicum, 13 from Vitis vinifera, 8 from Carica papaya, 13 from Arabidopsis thaliana, 12 from Ananas comosus, 25 from Brachypodium distachyon, 27 from Oryza sativa, 28 from Setaria italica, 28 from Setaria viridis, 27 from Zea mays, 29 from Sorghum bicolor, 30 from Saccharum spontaneum and 24 from Saccharum hybrid R570 [29]. The amino acid sequence of the 279 HAK/KUP/KT transporters from 15 representative plant species is provided in the supplementary data (Additional file 4).

Fig. 2
figure 2

Phylogeny and schematic diagram for intron/exon organization of HAK/KUP/KT genes from 15 plant species. a Clade II and clade III. b Clade I and clade IV

These HAK genes could be divided into four clades (I, II, III, IV) based on previously reported OsHAKs [17]. In A. trichopoda, the earliest diverging angiosperm, there were only 6 HAK genes, while in dicots and monocots, the number of HAKs ranged from 8 to 30 (Figs. 2 and 3), indicating that the ancient whole-genome duplication (WGD) contributed to the expansion of the HAK gene family in both dicots and monocots. Clade II and clade III included HAK genes from all 14 angiosperm genomes, indicating that the progenitors of these genes may have already existed prior to the split from angiosperms (Figs. 2 and 3). Clade I and clade IV mainly contained HAK genes from monocotyledons. Eighty-three HAK genes were identified in clade I, in which only one HAK gene was from A. comosus (Aco006685, homologous with SsHAK5) and Arabidopsis (AtHAK5), and the other 81 HAK genes were from all eight examined Poaceae species (Figs. 2 and 3). Twenty-nine HAKs were grouped into clade IV, and only 2 of them were from dicotyledons. These results indicated that the HAKs were unevenly distributed.

Fig. 3
figure 3

Phylogenetic relationships of HAK/KUP/KT families based on the current data for angiosperms

According to the Ks value in sorghum and sugarcane (Additional file 5), the divergence time of four clusters of HAKs was estimated. The median value of Ks was between 1.644 and 2.851, and its corresponding divergence time was between 134.8 and 233.7 Mya, indicating that the HAK was an ancient and divergent family. Furthermore, two pairs of duplicated SsHAKs (SsHAK5a/5b and SsHAK16a/16b) diverged at 18.94 and 58.14 Mya (Additional file 6). These results suggested that the SsHAK family is an ancient gene family with recent gene duplication events.

Exon/intron organization of the HAK family in S. spontaneum and other angiosperms

To investigate the structural characteristics and evolution of the HAK gene family, the exon/intron organization in HAKs was mapped to the phylogenic tree, and the gene features and patterns were analyzed (Fig. 2). The exon number in the HAK family of the 15 examined plant species ranged from 2 to 16, with an average of 8.4, and 217 out of 279 (77.8%) HAK genes possessed 8 to 10 exons (Additional files 7 and 8). This result suggested that the last common ancestor (LCA) of angiosperm HAK genes had 8 to 10 exons.

The exon number of SsHAKs varied from 2 to 12, and half of the SsHAKs possessed 8 or 9 exons. The pattern of SsHAK gene structure was similar to that of HAK gene structure from sorghum and maize in the same clade, suggesting that the HAK gene structure in the Panicoideae was relatively conserved. In clade I, the exon number in HAK genes varied from 2 to 12, which was the most variation among these 4 clades. Notably, the HAK genes in the subfamily with SsHAK22 had only 2 to 4 exons; however, the protein size remained consistent, which was likely due to the loss of introns. Clade II had the largest number of HAK genes, with 60 out of 98 HAKs possessing 9 exons and 5 out of 9 SsHAKs harboring 8 exons. SsHAK3/8/10 had one less exon than their orthologous genes in sorghum; the first exon in SsHAK13 and seventh exon in SsHAK24 were smaller than the corresponding exons in sorghum, and both resulted in shorter amino acid sequences in S. spontaneum (Table 1, Fig. 2). In clade III, the exon number was relatively conserved, with 61 out of 68 HAK genes possessing 8 to 10 exons, while the gene size varied greatly, mainly due to the different sizes of introns. The exon number in clade IV ranged from 2 to 8, with an average of 7, which was smaller than that in other clades. Notably, the HAK genes in the subfamily with SsHAK4 had only 2 to 5 exons, which was likely caused by intron loss during the process of evolution. The results indicated that HAKs underwent gene structure reconstruction under different evolutionary dynamics in S. spontaneum and other angiosperms in this study.

Expression analysis of HAK genes in Saccharum species

To study the expression profiles and potential functions of HAKs in Saccharum, we compared the gene expression patterns according to 4 sets of RNA-seq data: 1) different developmental stages and tissues; 2) a leaf gradient; 3) the circadian rhythm; and 4) treatment under low-potassium stress. The FPKM values of HAK1, HAK7 and HAK20b in YT55 at 0 h, 6 h, 12 h, 24 h, 48 h and 72 h under K+-starvation conditions were verified by RT-qPCR. The relative expression level was positively correlated with the FPKM value (R2 = 0.8419, Additional file 9), suggesting the reliability of gene expression based on the RNA-seq analysis.

Expression pattern of HAKs in different tissues at different stages

To study gene functional divergence among the Saccharum species, transcriptome profiles of HAKs between two Saccharum species, S. officinarum and S. spontaneum, were analyzed based on RNA-seq at three developmental stages (seedling, premature and mature stages) in five different tissues, 2 leaf (leaf and leaf roll) and 3 stalk (immature, maturing and mature) tissues (Fig. 4). Among the 30 HAK genes analyzed, 18 genes (HAK3/4/5a/5b/12/13/14/15/16a/16b/17/19a/19b/20a/20b/21/22/26) showed very low or undetectable expression levels in all examined tissues of the two Saccharum species. HAK1 and HAK2 had different expression patterns in the two Saccharum species. HAK1 had higher expression levels in S. spontaneum than in S. officinarum, and the expression level in leaves was higher than that in stems at three different stages. HAK2 had higher expression levels in S. officinarum than in S. spontaneum, and the expression level in stems was higher than that in leaves. HAK8 was mainly expressed in the upper stems, while the expression levels in the middle and lower stems were very low. HAK9 and HAK10 had higher expression levels in stems than in leaves. HAK18 was expressed in all examined tissues, with higher expression levels, especially in leaves at the seedling stage and in mature stems. Notably, HAK27 was highly expressed in leaves at all examined three stages, but the expression level in stems was very low or undetectable.

Fig. 4
figure 4

The expression pattern of HAK/KUP/KT genes based on FPKM in different tissues in different stages in S. officinarum and S. spontaneum

Expression pattern of HAKs across a leaf segment gradient

To further explore the functional divergence of HAK genes for photosynthesis in the source tissues, we studied the expression pattern of HAKs in continuously developing leaf segment gradients from S. officinarum and S. spontaneum (Fig. 5). Saccharum leaves were divided into four zones: the basal zone (sink tissue), transitional zone (undergoing sink-source transition), maturing zone and mature zone (fully differentiated zone with active photosynthesis), following the method described in maize [30]. Consistent with the expression pattern at different developmental stages, 18 HAK genes (HAK3/4/5a/5b/12/13/14/15/16a/16b/17/19a/19b/20a/20b/21/22/26) showed very low or undetected expression levels in all examined leaf segments, suggesting their limited roles in sugar transport (Fig. 5). HAK1 and HAK2 showed higher expression levels in the basal zone than in the other 3 zones. The expression level of HAK7 increased gradually from the base to the tip of the S. spontaneum leaf, while in S. officinarum, HAK7 displayed higher expression levels in the maturing zone than in the other 3 zones. The expression level of HAK8 decreased gradually from the base to the tip of the leaf in both S. officinarum and S. spontaneum. HAK9 showed different expression patterns in S. spontaneum and S. officinarum. In S. spontaneum, the expression level of HAK9 increased gradually from the basal zone to the maturing zone and then decreased in the mature zone. In S. spontaneum, the expression level of HAK9 decreased from the transition zone to the maturing zone and then increased in the mature zone, and the expression level was much higher in S. officinarum, suggesting gene functional divergence after the split of these two Saccharum species. HAK10 showed higher expression levels in the transition zone in S. spontaneum and higher expression levels in the mature zone in S. officinarum. HAK18 displayed higher expression levels in the maturing zone in both S. spontaneum and S. officinarum, while HAK23 showed higher expression levels in the basal zone in the two Saccharum species. HAK25 displayed higher expression levels in the maturing zone in S. officinarum but higher expression levels in the basal zone in S. spontaneum.

Fig. 5
figure 5

The expression pattern of HAK/KUP/KT genes based on FPKM across leaf gradients in S. officinarum and S. spontaneum

Expression pattern of HAKs during the circadian rhythm

Acting as an enzyme activator, potassium ions participate in a series of photosynthetic processes [31]. To analyze the expression pattern of HAKs during diurnal cycles, we investigated the transcriptome profiles of the mature leaves in the two Saccharum species at 2 h intervals over a 24 h period and at 4 h intervals over an additional 24 h. Consistent with the transcriptome profiles at different developmental stages and in the leaf segment gradient, 18 genes (HAK3/4/5a/5b/12/13/14/15/16a/16b/17/19a/19b/20a/20b/21/22/26) displayed very low or undetectable expression levels in the two Saccharum species, further supporting their limited roles in growth and development (Fig. 6). In addition, HAK8 and HAK24 also showed low expression levels over the two 24 h periods. HAK1, HAK2, HAK7, HAK18 and HAK27 showed higher expression levels in S. officinarum than in S. spontaneum, while HAK9 and HAK10 displayed higher expression levels in S. spontaneum than in S. officinarum. HAK1 and HAK2 had no diurnal expression pattern in the two Saccharum species. HAK7 displayed a higher expression level at night than in the daytime and showed the lowest expression level at noon in S. officinarum but showed no diurnal expression pattern in S. spontaneum. HAK10 displayed a higher expression level at night than in the daytime in S. spontaneum but showed no diurnal expression pattern in S. officinarum. HAK9 displayed a higher expression level at night than in the daytime in both Saccharum species. HAK18 and HAK27 displayed higher expression in the morning in the two Saccharum species. These findings suggested the functional divergence of the HAK genes in diurnal rhythms.

Fig. 6
figure 6

The expression pattern of HAK/KUP/KT genes based on FPKM during the diurnal cycles in S. officinarum and S. spontaneum

Expression pattern of HAKs under K+-deficiency stress

To investigate the functional divergence of HAK genes in response to low-potassium stress in sugarcane, we studied the expression profiles of HAKs in roots from the Saccharum hybrid variety YT55 at 0 h, 6 h, 12 h, 24 h, 48 h and 72 h under low-K+ stress (Fig. 7). Among the 30 HAK genes analyzed, 14 genes (HAK3/4/5a/5b/11/13/16a/16b/19a/19b/20a/22/26/27) showed very low or undetectable expression levels before and after exposure to low-K+ stress. Notably, HAK1 showed strong induction in roots under low-K+ conditions, reached the highest level at 24 h, and then decreased subsequently at 48 h and 72 h. HAK21 was strongly induced after exposure to low-K+ stress within 12 h but was subsequently downregulated to a low expression level. HAK20b was downregulated within 12 h and then upregulated to the highest level at 72 h. HAK7, HAK10, HAK18 and HAK24 were downregulated after exposure to low-K+ stress. Other HAKs, such as HAK12/14/15/25, were constitutively expressed.

Fig. 7
figure 7

a The expression pattern of HAK/KUP/KT genes in Saccharum hybrid YT55 under low-K+ stress conditions based on FPKM values. b The relative expression level detected by RT-qPCR

Functional complementation validation of SsHAK1 and SsHAK21 in the yeast mutant strain R5421

SsHAK1 and SsHAK21 were selected for complementary validation in yeast because they were both induced in response to low-K+ stress. The transformed yeast strain carrying only the empty vector pYES2.0 was used as a control. There were no obvious growth differences between yeast transformed with pYES2.0 and yeast transformed with pYES2.0-SsHAK1 or pYES2.0-SsHAK21 in SC/−ura medium containing 100 mM KCl (Fig. 8). However, when the KCl concentration decreased to 10 mM, the growth of yeast transformed with SsHAK1 and SsHAK21 was better than that of yeast transformed with the empty vector. When the KCl concentration decreased to 1 mM, the growth of yeast transformed with the empty vector was significantly inhibited, while the growth of yeast transformed with SsHAK1 or SsHAK21 was almost unaffected (Fig. 8). These results suggested that both SsHAK1 and SsHAK21 could recover the K+ absorption function in the yeast mutant strain R5421, indicating that they had potassium transporter activity.

Fig. 8
figure 8

Phenotypic identification of the yeast mutant strain R5421 transformed with SsHAK1 or SsHAK21

Discussion

The HAK/KUP/KT family of potassium transporters have been widely reported to be associated with K+ transport across membranes in plants. They also play vital roles in response to salt and drought stress, as well as morphogenesis of root and shoot [7]. However, genome-wide analysis of the HAK/KUP/KT gene family has not been conducted in Saccharum due to its complex genetic background. The recently released S. spontaneum genome allowed us to identify 30 HAK genes from S. spontaneum. In addition, 248 HAK genes from 13 other representative plant species and an outgroup were used to construct a phylogenetic tree and study the evolution of HAK genes in Saccharum. Furthermore, expression analysis based on RNA-seq revealed spatiotemporal expression and functional divergence in the HAK family, which provides valuable information and robust candidate genes for future functional analysis.

Evolution of the HAK gene family in Saccharum and representative angiosperms

WGD or polyploidization, gene loss and diploidization are considered important evolutionary forces in plants [32, 33]. Angiosperms, pancore eudicots and monocots originated from ε, γ and σ WGD events, which have been revealed by a rigorous phylogenomic approach [33]. A recent study showed that pineapple had one fewer ancient ρ WGD events than other gramineous plants [34]. A. trichopoda is the earliest known angiosperm to have evolved separately from other angiosperms and has attracted much attention from botanists. In this study, 279 HAKs from 15 plant species representing major WGD events in angiosperms, together with the WGD information, allowed us to study HAK gene evolution. HAKs from different plant species could be divided into four clades in in descending order of duplications: clade IV, clade I, clade III and clade II. Based on the estimated divergence time among the 4 clades of the SsHAK gene family (134.8 to 233.7 Mya, Additional file 5), the SsHAK family in angiosperms probably occurred before the σ WGD event in angiosperms (approximately 130 Mya) and after the ε WGD event (approximately 220 Mya) [33].

The number of HAKs in the four clades varied greatly (from 29 to 98, Fig. 3), which is consistent with a previous study in which HAKs were unevenly distributed in different clades among angiosperms [35]. In clade I, only one HAK gene member was from A. comosus and Arabidopsis, while in Poaceae species, the HAK number ranged from 6 to 13. This result indicated that WGD or recent gene duplication contributed greatly to the expansion of HAKs. SsHAK5a/5b, SsHAK16a/16b, and SsHAK19a/19b were from tandem duplications, while SsHAK20a/20b may have originated from a transposed duplication. The LCAs of SsHAK5 and SsHAK18 (in clade III) may have occurred before the split of monocotyledonous and dicotyledonous plants. HAK5 was speculated to be lost in other dicotyledons except for Arabidopsis, which may be due to the gene functional redundancy of the HAK family. HAK18 was retained in all monocotyledonous and dicotyledonous plants, showing its functional constraint for the HAK family, and the expression profile analysis of HAK18 also confirmed this.

In clade II and clade III, SsHAK2 and SsHAK7 were retained from the ε WGD event, and in dicotyledons, these two orthologous genes were lost. SsHAK3 and SsHAK13 originated after A. trichopoda had evolved separately from other angiosperms. SsHAK8, SsHAK9 and SsHAK10 were assumed to be retained from the ε WGD event; SsHAK11, SsHAK12, SsHAK15, SsHAK24 and SsHAK25 were retained from the σ WGD event, as only monocotyledons contained these genes. SsHAK14 and SsHAK23 were assumed to be retained from the ε WGD event, but HAK14 was probably lost in dicotyledons. Clade IV contained the lowest number of HAKs. SsHAK4 and SsHAK17 originated before the split of monocotyledons and dicotyledons and after the split of A. trichopoda from angiosperms. The LCA of SsHAK26 originated after the split of the Gramineae and pineapple.

The HAK gene family in plants exhibited a less conserved exon/intron structure. The exon number in Saccharum ranged from 2 to 12 (Fig. 1, Additional file 7), and the variation range in Saccharum was larger than that in rice [17], maize [19] and wheat [36]. Three types of mechanisms, exon/intron gain/loss, exonization/pseudoexonization and insertion/deletion, mainly led to exon-intron structure differences in paralogous or orthologous genes [37]. Although the gene structure of SsHAKs changed greatly, the protein size was relatively conserved, suggesting that exon-intron structure differences in SsHAKs were mainly caused by intron gain/loss. Clade I and clade IV belong to the older HAK family in Saccharum, so the HAKs in these two clades were speculated to have more intron gain/loss events based on the ‘introns-early’ theory during the lengthy evolutionary process [38, 39]. The results in this study also support this view because the variation in exon number in clade I and clade IV was much greater than that in clade II and clade III.

Gene expression and functional divergence of HAKs in Saccharum

The transcriptional regulation of K+ transporters is a common mechanism by which plant species respond to low-K+ stress [8], and expression pattern analysis can provide insight into the potential functions of the HAK gene family. In this study, we found that most HAK genes in clade I and clade IV showed low or undetectable expression levels across all examined samples. Most HAK genes in clade II and clade III were strongly expressed in all tested tissues. These results were consistent with the results of previous studies on HAK genes in rice [17], Arabidopsis [25] and wheat [36]. Five OsHAK genes (OsHAK2/10/15/23/25) from clades II and III were expressed in all examined tissues of three different genotypes [17]. In Arabidopsis, 12 out of 13 HAK/KUP/KT genes were from clades II and III, most of which were expressed in the roots, leaves, siliques and flowers [25]. Similarly, most TaHAKs in wheat belonging to clades II and III were constitutively expressed in all tissues [36].

Low-K+ stress tends to induce the upregulated expression of K+ transporter genes [40]. Previous studies demonstrated that the expression of OsHAK1 in rice [20], TaHAK1 in wheat [36] and PbrHAK1 in pear [41] was induced by K+ starvation. In this study, the expression level of SsHAK1 increased rapidly under low-K+ stress, and this result is in good agreement with previous studies. Notably, SsHAK21 was substantially upregulated after a short period of K+-starvation treatment and then rapidly downregulated (transient activation), indicating that SsHAK21 was involved in the low-K+ stress response in sugarcane. Similar results were found in rice, as OsHAK21 functions in the maintenance of ion homeostasis and tolerance to salt stress [42]. SsHAK1, SsHAK17 and SsHAK21 displayed upregulated expression, suggesting that they may play important roles in maintaining normal growth and mediating potassium acquisition under K+ deficiency. In addition, nearly half of the SsHAK genes were not expressed or had very low levels of expression in all tested tissues at all stages or even under low-K+ stress, which may be caused by the gene functional redundancy due to WGD events in sugarcane.

The root system acquires K+ from the soil solution, and then K+ is transported among compartments within cells and from the roots to the shoots. A schematic model was proposed based on the expression profiles of the 30 SsHAK genes to illustrate the spatial and temporal gene expression in plant tissues and root hair cells of sugarcane (Fig. 9). HAK7/9/18 were mainly expressed in the tissues of maturing and mature stems and leaves, indicating their important roles in K+ transport in these tissues. HAK7/9/18/25 also showed a circadian rhythm expression pattern, suggesting that these genes were regulated by sunlight. Low-K+ stress induced the upregulation of the transcriptional expression of HAK genes. In Arabidopsis, transcription factors, such as DDF2, JLO, ARF2, RAP2.11, TFII_A, and bHLH121, directly bind the promoter of AtHAK5 to induce its expression and increase tolerance to low-K+ and salt stress [26]. In this study, the expression of HAK1 and HAK21 was greatly upregulated, which may also be positively regulated by transcription factors (TFs), such as DDF2 and JLO, and further experiments, such as yeast one-hybrid assays, can be used to screen the TFs. AtHAK5 and its homologs from pepper and tomato can be activated by the CIPK23 (CBL-interacting protein kinase 23)/CBL1 (calcineurin B-like protein) complex [27]. In rice, OsHAK1/19/20 can be phosphorylated by a receptor-like protein kinase, RUPO (ruptured pollen tube) [43]. In this study, the CBL-CIPK complex and the receptor-like kinase RUPO may also act as regulators of high-affinity potassium transporters, such as HAK1, via phosphorylation-dependent interactions.

Fig. 9
figure 9

Schematic models for the roles of HAKs based on gene expression profiles in sugarcane. In the maturing and mature zones of the leaves and stems, HAK7/9/18 were the main expressed genes. Moreover, these genes also presented a diurnal expression pattern. HAK25 was mainly expressed in the maturing and mature zone of leaf tissues, while HAK2 was mainly expressed in the stem. Low-K+ stress induced the upregulation of the expression of HAK1 and HAK21. Transcription factors such as DDF2 and JLO may directly bind to the promoters of HAK1/21 to induce gene expression and subsequently promote HAK transporters, such as HAK1 and HAK 21, to acquire K+ in roots. HAK1 may be phosphorylated and activated by the CBL1-CIPK23 complex or receptor-like kinase, RUPO (ruptured pollen tube). The K+ concentration in the vacuole is highly varied to maintain cellular K+ homeostasis. Some HAK transporters, such as HAK10, located in the tonoplast of vacuoles may play a role in regulating the K+ concentration in vacuoles

Conclusions

In this study, 30 HAK (high-affinity K+ transporter) genes were identified through comparative genomics analyses of sugarcane. Evolutionary analysis revealed that both ancient whole-genome duplication (WGD) and recent gene duplication contributed to the expansion of the gene family, and purifying selection was the main force driving evolution. HAK/KUP/KT genes were accompanied by intron gain/loss in the process of evolution. Expression analysis based on RNA-seq under low-K+ stress and at different developmental stages revealed spatiotemporal expression and functional divergence in the HAK/KUP/KT gene family. Yeast functional complementation analysis showed that SsHAK1 and SsHAK21 mediated K+ transport under low-K+ stress. Altogether, these results provide valuable information and robust candidate genes for future functional analyses for the genetic improvement of potassium-utilization efficiency in sugarcane.

Methods

Plant materials

Two Saccharum species, LA-Purple (S. officinarum, 2n = 8x = 80, originated in the USA and was introduced into China; the plants were provided by Zhang’s laboratory at Fujian Agriculture and Forestry University) and SES-208 (S. spontaneum, 2n = 8x = 64, originated in the USA and was introduced into China; the plants were provided by Zhang’s laboratory in Fujian Agriculture and Forestry University), were cultivated at Fujian Agricultural and Forestry University (Fuzhou, 119°16′48″E, 26°4′48″N, Fujian, China) and sampled for gene expression pattern analysis.

The K+ uptake-deficient yeast mutant strain R5421 (ura3–52 his3Δ200 leu2Δ1 trp1Δ1 ade2 trk1Δ::HIS3 trk2Δ::HIS3) was provided by Professor Guohua Xu from Nanjing Agricultural University. R5421 cannot grow normally when the external potassium concentration is below 10 mM. E. coli DH5α competent cells and the expression vector pYES2.0 were purchased from TaKaRa Biotechnology Co., Ltd. (Dalian, China).

For expression pattern analysis at different developmental stages, tissue samples including leaf roll, leaf (fully expanded leaf), top immature internode, premature internode and mature internode were collected from premature plants (9-month-old plants) and mature plants (12-month-old plants). The sugarcane internodes were numbered from top to bottom. Leaf and stem tissues in the seedling stage were collected from 35-day-old plants as previously described [44, 45].

For expression pattern analysis of the leaf gradient, the two Saccharum species were grown in a greenhouse with light intensities of 350 μmol/m2/sec, 14:10 L/D, 30 °C L/22 °C D and 60% relative humidity. The second leaf of 15-day-old LA-Purple and 11-day-old SES208 after planting at 3 h into the light period and samples preparation method was as described by Li et al. [30].

For expression pattern analysis of the diurnal cycle, leaves of the mature plants of LA-Purple and SES208 were sampled consecutively 12 times with 2 h intervals from 6:00 a.m. on March 2, 2017, then sampled consecutively 7 times with 4 h intervals from 6:00 a.m. on March 3, 2017. The sunrise and sunset times on March 2, 2017 in Fuzhou were 6:25 a.m. and 6:05 p.m. respectively. Tissue collection was performed following a previously described method [34].

For expression pattern analysis under low-potassium stress, Saccharum hybrid variety YT55 (this variety was bred by Guangzhou Sugarcane Industry Research Institute and was planted in breeding bases for sugarcane in Wengyuan, Guangdong Province) was cultured at a normal potassium level (3.0 mmol /L) for 20 days in a greenhouse and then transferred to the K+-deficient nutrient solution (0.1 mmol /L) for starvation treatment. Mixed samples of roots from 6 plants in a pot (a biological replicate and three biological replicates in total were collected) were collected at 0 h, 6 h, 12 h, 24 h, 48 h and 72 h after starvation and stored in liquid nitrogen for total RNA isolation.

Homology search analysis

According to previous reports, the protein sequences of 13, 27 and 27 HAK/KUP/ KT gene families identified in Arabidopsis thaliana, Oryza sativa and Zea mays [17,18,19] were obtained from Phytozome V12.1 (https://phytozome.jgi.doe.gov/pz/portal.html). With these protein sequences as queries, putative members of the HAK/KUP/ KT gene family were searched using the BLASTP program in 14 representative angiosperm genomes, 9 monocotyledons (Saccharum hybrid R570 [29], Saccharum spontaneum, Sorghum bicolor, Zea mays, Setaria viridis, Setaria italica, Oryza sativa, Brachypodium distachyon and Ananas comosus), 4 dicotyledons (Arabidopsis thaliana, Carica papaya, Vitis vinifera, and Solanum lycopersicum) and Amborella trichopoda. Sequences with an e-value <1e− 10 were selected as HAK/KUP/KT candidates. Then, the identified HAK/KUP/KT proteins were subjected to conserved domain validation with the PFAM (https://pfam.xfam.org) and CDD (https://www.ncbi.nlm.nih.gov/Structure/bwrpsb/bwrpsb.cgi) databases. In addition, a HAK gene from Chlamydomonas reinhardtii was selected as the outgroup.

Sequence and phylogenetic analyses

Isoelectric points (pI) and relative molecular weight of the HAK/KUP/KT proteins were predicted by ExPASy (https://web.expasy.org/compute_pi/). The exon-intron structures were assessed with TBtools [46]. TMHHM Server v.2.0 (http://www.cbs.dtu.dk/ services/TMHMM/) was used to predict the transmembrane domains of the HAK/KUP/KT proteins. Subcellular locations of the HAK/KUP/KT proteins were predicted by WoLF PSORT (https://www.genscript.com/wolf-psort.html).

The evolutionary history of 14 representative angiosperms was inferred by the neighbor-jointing (NJ) method [47]. Based on the protein sequence alignment, the phylogenetic tree of the HAK/KUP/KT gene family was constructed using NJ methods. The construction of the NJ tree was performed using MEGA7 [48] with the “pair deletion” and “Poisson correction” models. The reliability of the internal branches of the tree was evaluated by a bootstrap test (1000 replicates), and the percentages are shown next to the branches.

The nonsynonymous substitution ratios (Ka), synonymous substitution ratios (Ks) and Ka/Ks ratios of the 30 pairs HAK/KUP/KT orthologous genes from sorghum and sugarcane were calculated by the Easy_KaKs calculation program (https://github.com/tangerzhang/FAFU-cgb/tree/master/easy_KaKs). Fisher’s exact test for small samples was applied to verify the validity of Ka and Ks calculated by this method [49]. The divergence time (T) was calculated as T = Ks/ (2 × 6.1 × 10− 9) × 10− 6 Mya [50].

Analysis of the expression profiling of HAKs in Saccharum based on RNA-seq

RNA preparation, cDNA libraries construction and RNA-seq libraries sequencing were performed as previously described [51, 52]. Raw data were aligned to available S. spontaneum AP85–441 reference gene models using Trinity (https://github.com/trinityrnaseq/trinityrnaseq/wiki). Fragments per kilobase per million mapped fragments (FPKM) values were calculated to represent gene expression levels as previously described [51, 52].

Validation of HAK gene expression levels by RT-qPCR

The expression level of three HAK genes (HAK1, HAK7 and HAK20b) in the roots of Saccharum hybrid variety YT55 at 6 time points (0 h, 6 h, 12 h, 24 h, 48 h and 72 h) under K+-starvation conditions was validated by RT-qPCR, to normalize the expression levels, 2 constitutively expressed genes, the eukaryotic elongation factor 1a (eEF-1a) and actin were used as reference genes, each sample had 3 biological replicates and 3 technical replicates. (Additional file 10). The reaction program of reverse transcription, real-time PCR and the relative expression levels calculation were carried out as Wang et al. described [52].

Yeast expression vector construction and function complementation experiment of SsHAK1 and SsHAK21

Primer Premier 5 was used to design primers (Additional file 11), and the synthesized cDNA from RNA of YT55 after 12 h of low-potassium stress treatment was used as a template to amplify SsHAK1 and SsHAK21. The amplified products were recovered from the gel and ligated to the expression vector pYES2.0 with In-Fusion enzyme (TaKaRa Biotechnology Co., Ltd., Dalian, China). The ligation products were transformed into E. coli competent DH5α cells. Positive monoclonal clones were selected and verified by sequencing, and then the plasmids were extracted for subsequent yeast transformation. Competent cells of yeast mutant strain R5421 were prepared with the S.c. EasyComp™ Transformation Kit (Invitrogen, Carlsbad, CA, USA) and transformed. Yeast strains with the empty vector and target genes were isolated and then used for gradient dilution and inoculated in SC/−ura medium with 100 mM, 5 mM and 0 mM KCl. The results were observed after 3–5 days of culture at 30 °C.

Availability of data and materials

The datasets supporting the conclusions of this article are included in the article and its additional files.

Abbreviations

bHLH121:

Basic helix-loop-helix 121

CBL:

Calcineurin B-like protein

CIPK:

CBL-interacting protein kinase

DDF2:

Dwarf and delayed flowering 2

FPKM:

Fragments per kilobase per million mapped fragments

HAK/KUP/KT:

High-affinity K+ transporter/K+ uptake permease/K+ transporter

JLO:

Jagged lateral organs

Ka:

Nonsynonymous substitution ratio

Ks:

Synonymous substitution ratio

LCA:

Last common ancestor

RT-qPCR:

Reverse transcription-quantitative PCR

TF:

Transcription factor

TFII_A:

Transcription initiation factor II_A gamma chain

WGD:

Whole-genome duplication

References

  1. Leigh RA, Wyn Jones RG. A hypothesis relating critical potassium concentrations for growth to the distribution and function of this ion in the plant cell. New Phytol. 1984;97:1–13.

    Article  CAS  Google Scholar 

  2. Ashley MK, Grant M, Grabov A. Plant responses to potassium deficiencies: a role for potassium transport proteins. J Exp Bot. 2006;57(2):425–36.

    Article  CAS  PubMed  Google Scholar 

  3. Coale FJ, Izuno FT, Bottcher AB. Nutrient accumulation and removal by sugarcane grown on Everglades Histosols. Agron J. 1993;85:310–5.

    Article  CAS  Google Scholar 

  4. Wood RA. The roles of nitrogen, phosphorus and potassium in the production of sugarcane in South Africa. Fertilizer Res. 1990;26:89–98.

    Article  CAS  Google Scholar 

  5. White PJ. Improving potassium acquisition and utilisation by crop plants. J Plant Nutr Soil Sci. 2013;176(3):305–16.

    Article  CAS  Google Scholar 

  6. Epstein E, Rains DW, Elzam OE. Resolution of dual mechanisms of potassium absorption by barley roots. Proc Natl Acad Sci U S A. 1963;49(5):684–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Li W, Xu G, Alli A, Yu L. Plant HAK/KUP/KT K(+) transporters: function and regulation. Semin Cell Dev Biol. 2018;74:133–41.

    Article  CAS  PubMed  Google Scholar 

  8. Wang Y, Wu WH. Potassium transport and signaling in higher plants. Annu Rev Plant Biol. 2013;64:451–76.

    Article  CAS  PubMed  Google Scholar 

  9. Corratge-Faillie C, Jabnoune M, Zimmermann S, Very AA, Fizames C, Sentenac H. Potassium and sodium transport in non-animal cells: the Trk/Ktr/HKT transporter family. Cell Mol Life Sci. 2010;67(15):2511–32.

    Article  CAS  PubMed  Google Scholar 

  10. Schleyer M, Bakker EP. Nucleotide sequence and 3′-end deletion studies indicate that the K(+)-uptake protein kup from Escherichia coli is composed of a hydrophobic core linked to a large and partially essential hydrophilic C terminus. J Bacteriol. 1993;175(21):6925–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Fu HH, Luan S. AtKuP1: a dual-affinity K+ transporter from Arabidopsis. Plant Cell. 1998;10(1):63–73.

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Santa-Maria GE, Rubio F, Dubcovsky J, Rodriguez-Navarro A. The HAK1 gene of barley is a member of a large gene family and encodes a high-affinity potassium transporter. Plant Cell. 1997;9(12):2281–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Banuelos MA, Garciadeblas B, Cubero B, Rodriguez-Navarro A. Inventory and functional characterization of the HAK potassium transporters of rice. Plant Physiol. 2002;130(2):784–95.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Gierth M, Maser P, Schroeder JI. The potassium transporter AtHAK5 functions in K(+) deprivation-induced high-affinity K(+) uptake and AKT1 K(+) channel contribution to K(+) uptake kinetics in Arabidopsis roots. Plant Physiol. 2005;137(3):1105–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Kim EJ, Kwak JM, Uozumi N, Schroeder JI. AtKUP1: an Arabidopsis gene encoding high-affinity potassium transport activity. Plant Cell. 1998;10(1):51–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Martinez-Cordero MA, Martinez V, Rubio F. Cloning and functional characterization of the high-affinity K+ transporter HAK1 of pepper. Plant Mol Biol. 2004;56(3):413–21.

    Article  CAS  PubMed  Google Scholar 

  17. Gupta M, Qiu X, Wang L, Xie W, Zhang C, Xiong L, Lian X, Zhang Q. KT/HAK/KUP potassium transporters gene family and their whole-life cycle expression profile in rice (Oryza sativa). Mol Gen Genomics. 2008;280(5):437–52.

    Article  CAS  Google Scholar 

  18. Mäser P, Thomine S, Schroeder JI, Ward JM, Hirschi K, Sze H, et al. Phylogenetic relationships within cation transporter families of Arabidopsis. Plant Physiol. 2001;126:1646–67.

    Article  PubMed  PubMed Central  Google Scholar 

  19. Zhang Z, Zhang J, Chen Y, Li R, Wang H, Wei J. Genome-wide analysis and identification of HAK potassium transporter gene family in maize (Zea mays L.). Mol Biol Rep. 2012;39(8):8465–73.

    Article  CAS  PubMed  Google Scholar 

  20. Chen G, Hu Q, Luo L, Yang T, Zhang S, Hu Y, Yu L, Xu G. Rice potassium transporter OsHAK1 is essential for maintaining potassium-mediated growth and functions in salt tolerance over low and high potassium concentration ranges. Plant Cell Environ. 2015;38(12):2747–65.

    Article  CAS  PubMed  Google Scholar 

  21. Yang T, Zhang S, Hu Y, Wu F, Hu Q, Chen G, Cai J, Wu T, Moran N, Yu L, et al. The role of a potassium transporter OsHAK5 in potassium acquisition and transport from roots to shoots in rice at low potassium supply levels. Plant Physiol. 2014;166(2):945–59.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Nieves-Cordones M, Miller AJ, Aleman F, Martinez V, Rubio F. A putative role for the plasma membrane potential in the control of the expression of the gene encoding the tomato high-affinity potassium transporter HAK5. Plant Mol Biol. 2008;68(6):521–32.

    Article  CAS  PubMed  Google Scholar 

  23. Rubio F, Nieves-Cordones M, Aleman F, Martinez V. Relative contribution of AtHAK5 and AtAKT1 to K+ uptake in the high-affinity range of concentrations. Physiol Plant. 2008;134(4):598–608.

    Article  CAS  PubMed  Google Scholar 

  24. Su H, Golldack D, Zhao C, Bohnert HJ. The expression of HAK-type K(+) transporters is regulated in response to salinity stress in common ice plant. Plant Physiol. 2002;129(4):1482–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Ahn SJ, Shin R, Schachtman DP. Expression of KT/KUP genes in Arabidopsis and the role of root hairs in K+ uptake. Plant Physiol. 2004;134(3):1135–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Hong JP, Takeshi Y, Kondou Y, Schachtman DP, Matsui M, Shin R. Identification and characterization of transcription factors regulating Arabidopsis HAK5. Plant Cell Physiol. 2013;54(9):1478–90.

    Article  CAS  PubMed  Google Scholar 

  27. Ragel P, Rodenas R, Garcia-Martin E, Andres Z, Villalta I, Nieves-Cordones M, Rivero RM, Martinez V, Pardo JM, Quintero FJ, et al. The CBL-interacting protein kinase CIPK23 regulates HAK5-mediated high-affinity K+ uptake in Arabidopsis roots. Plant Physiol. 2015;169(4):2863–73.

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Zhang J, Zhang X, Tang H, Zhang Q, Hua X, Ma X, Zhu F, Jones T, Zhu X, Bowers J, et al. Allele-defined genome of the autopolyploid sugarcane Saccharum spontaneum L. Nat Genet. 2018;50(11):1565–73.

    Article  CAS  PubMed  Google Scholar 

  29. Garsmeur O, Droc G, Antonise R, Grimwood J, Potier B, Aitken K, Jenkins J, Martin G, Charron C, Hervouet C, et al. A mosaic monoploid reference sequence for the highly complex genome of sugarcane. Nat Commun. 2018;9(1):2638.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Li P, Ponnala L, Gandotra N, Wang L, Si Y, Tausta SL, Kebrom TH, Provart N, Patel R, Myers CR, et al. The developmental dynamics of the maize leaf transcriptome. Nat Genet. 2010;42(12):1060–7.

    Article  CAS  PubMed  Google Scholar 

  31. Lu Z, Xie K, Pan Y, Ren T, Lu J, Wang M, Shen Q, Guo S. Potassium mediates coordination of leaf photosynthesis and hydraulic conductance by modifications of leaf anatomy. Plant Cell Environ. 2019. https://doi.org/10.1111/pce.13553.

  32. Edger PP, Pires JC. Gene and genome duplications: the impact of dosage-sensitivity on the fate of nuclear genes. Chromosom Res. 2009;17(5):699–717.

    Article  CAS  Google Scholar 

  33. Jiao Y, Wickett NJ, Ayyampalayam S, Chanderbali AS, Landherr L, Ralph PE, Tomsho LP, Hu Y, Liang H, Soltis PS, et al. Ancestral polyploidy in seed plants and angiosperms. Nature. 2011;473(7345):97–100.

    Article  CAS  PubMed  Google Scholar 

  34. Ming R, VanBuren R, Wai CM, Tang H, Schatz MC, Bowers JE, Lyons E, Wang ML, Chen J, Biggers E, et al. The pineapple genome and the evolution of CAM photosynthesis. Nat Genet. 2015;47(12):1435–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Nieves-Cordones M, Rodenas R, Chavanieu A, Rivero RM, Martinez V, Gaillard I, Rubio F. Uneven HAK/KUP/KT protein diversity among angiosperms: species distribution and perspectives. Front Plant Sci. 2016;7:127.

    PubMed  PubMed Central  Google Scholar 

  36. Cheng X, Liu X, Mao W, Zhang X, Chen S, Zhan K, Bi H, Xu H. Genome-wide identification and analysis of HAK/KUP/KT potassium transporters gene family in wheat (Triticum aestivum L.). Int J Mol Sci. 2018;19(12):3969.

    Article  PubMed Central  Google Scholar 

  37. Xu G, Guo C, Shan H, Kong H. Divergence of duplicate genes in exon-intron structure. Proc Natl Acad Sci U S A. 2012;109(4):1187–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Jeffares DC, Mourier T, Penny D. The biology of intron gain and loss. Trends Genet. 2006;22(1):16–22.

    Article  CAS  PubMed  Google Scholar 

  39. Rogozin IB, Sverdlov AV, Babenko VN, Koonin EV. Analysis of evolution of exon-intron structure of eukaryotic genes. Brief Bioinform. 2005;6(2):118–34.

    Article  CAS  PubMed  Google Scholar 

  40. Wang Y, Wu WH. Regulation of potassium transport and signaling in plants. Curr Opin Plant Biol. 2017;39:123–8.

    Article  CAS  PubMed  Google Scholar 

  41. Wang Y, Lu J, Chen D, Zhang J, Qi K, Cheng R, Zhang H, Zhang S. Genome-wide identification, evolution, and expression analysis of the KT/HAK/KUP family in pear. Genome. 2018;61(10):755–65.

    Article  CAS  PubMed  Google Scholar 

  42. Shen Y, Shen L, Shen Z, Jing W, Ge H, Zhao J, Zhang W. The potassium transporter OsHAK21 functions in the maintenance of ion homeostasis and tolerance to salt stress in rice. Plant Cell Environ. 2015;38(12):2766–79.

    Article  CAS  PubMed  Google Scholar 

  43. Liu L, Zheng C, Kuang B, Wei L, Yan L, Wang T. Receptor-like kinase RUPO interacts with potassium transporters to regulate pollen tube growth and integrity in rice. PLoS Genet. 2016;12(7):e1006085.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  44. Chen Y, Zhang Q, Hu W, Zhang X, Wang L, Hua X, Yu Q, Ming R, Zhang J. Evolution and expression of the fructokinase gene family in Saccharum. BMC Genomics. 2017;18(1):197.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  45. Zhang Q, Hu W, Zhu F, Wang L, Yu Q, Ming R, Zhang J. Structure, phylogeny, allelic haplotypes and expression of sucrose transporter gene families in Saccharum. BMC Genomics. 2016;17:88.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Chen CJ, Xia R, Chen H, He YH. TBtools, a toolkit for biologists integrating various HTS-data handling tools with a user-friendly interface. BioRxiv Preprint, 2018.

    Book  Google Scholar 

  47. Saitou N, Nei M. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evol. 1987;4(4):406–25.

    CAS  PubMed  Google Scholar 

  48. Kumar S, Stecher G, Tamura K. MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol Biol Evol. 2016;33(7):1870–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Graham JGU. Fisher’s exact test. J R Stat Soc Ser A (Stat Soc). 1992;155(3):395–402.

    Article  Google Scholar 

  50. Lynch M, Conery JS. The evolutionary fate and consequences of duplicate genes. Science. 2000;290(5494):1151–5.

    Article  CAS  PubMed  Google Scholar 

  51. Hu W, Hua X, Zhang Q, Wang J, Shen Q, Zhang X, Wang K, Yu Q, Lin YR, Ming R, et al. New insights into the evolution and functional divergence of the SWEET family in Saccharum based on comparative genomics. BMC Plant Biol. 2018;18(1):270.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Wang Y, Hua X, Xu J, Chen Z, Fan T, Zeng Z, Wang H, Hour AL, Yu Q, Ming R, et al. Comparative genomics revealed the gene evolution and functional divergence of magnesium transporter families in Saccharum. BMC Genomics. 2019;20(1):83.

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgments

We kindly thank the Center for Genomics and Biotechnology, Haixia Institute of Science and Technology, Fujian Agriculture and Forestry University for providing access to Saccharum data. The authors gratefully acknowledge Professor Guohua Xu from Nanjing Agricultural University for providing the yeast mutant strain.

Funding

This work was supported by funding from the GDAS’ Project of Science and Technology Development (2019GDASYL-0103028, used to purchase reagents and consumables for RT-qPCR of HAK1, HAK7 and HAK20b); special funds from the Guangdong Academy of Science (2019GDASYL-0104013 and 2017GDASCX-0105, used to pay for RNA-seq of YT55 under low-potassium stress); the Guangdong Provincial Team of Technical System Innovation for Sugarcane Sisal Hemp Industry (2019KJ104–04, used to pay for technicians to plant sugarcane YT55); the Science and Technology Planting Project of Guangdong Province, China (2014B070705002, used for travel expenses between Fuzhou and Guangzhou to perform yeast complementary experiment); and the China Agricultural Research System (CARS201707, used for article-processing charges of this manuscript).

Author information

Authors and Affiliations

Authors

Contributions

XF, JZ and YQ conceived the study and designed the experiments. XF, YW, NZ, ZW, QZ, JW, XW, LW and JZ carried out the experiments and analyzed the data. XF wrote the manuscript. All authors read and approved the final paper.

Corresponding author

Correspondence to Yongwen Qi.

Ethics declarations

Ethics approval and consent to participate

Not applicable

Consent for publication

Not applicable

Competing interests

The authors declare that they have no competing interests.

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.

The HAK gene alleles in Saccharum spontaneum.

Additional file 2.

Similarity between HAK proteins in sugarcane calculated by NCBI BLASTP.

Additional file 3.

Phylogenetic relationships among the KT/HAK/KUP gene families from 15 representative plant species.

Additional file 4.

Amino acid sequence of 279 HAK/KUP/KT transporters from 15 representative plant species.

Additional file 5.

Divergence time among the 4 clades of the HAK family in Sorghum bicolor and Saccharum spontaneum.

Additional file 6.

Divergence between paralogous SsHAK gene pairs in Saccharum spontaneum.

Additional file 7.

Statistics of exon number in each HAK.

Additional file 8.

The proportion of different numbers of exons in all HAKs from 15 plant species.

Additional file 9.

Correlation coefficient between RNA-seq data and RT-qPCR of HAK1, HAK7 and HAK20b.

Additional file 10.

The primers for the RT-qPCR verification of four HAK genes in Saccharum hybrid YT55.

Additional file 11.

The primers used to clone SsHAK1 and SsHAK21 and construct the yeast expression vector.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. 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.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Feng, X., Wang, Y., Zhang, N. et al. Genome-wide systematic characterization of the HAK/KUP/KT gene family and its expression profile during plant growth and in response to low-K+ stress in Saccharum. BMC Plant Biol 20, 20 (2020). https://doi.org/10.1186/s12870-019-2227-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12870-019-2227-7

Keywords