Skip to main content

A PIP-mediated osmotic stress signaling cascade plays a positive role in the salt tolerance of sugarcane

Abstract

Background

Plasma membrane intrinsic proteins (PIPs) are plant channel proteins involved in water deficit and salinity tolerance. PIPs play a major role in plant cell water balance and responses to salt stress. Although sugarcane is prone to high salt stress, there is no report on PIPs in sugarcane.

Results

In the present study, eight PIP family genes, termed ScPIP1–1, ScPIP1–2, ScPIP1–3, ScPIP1–4, ScPIP2–1, ScPIP2–2, ScPIP2–4 and ScPIP2–5, were obtained based on the sugarcane transcriptome database. Then, ScPIP2–1 in sugarcane was cloned and characterized. Confocal microscopy observation indicated that ScPIP2–1 was located in the plasma membrane and cytoplasm. A yeast two-hybridization experiment revealed that ScPIP2–1 does not have transcriptional activity. Real time quantitative PCR (RT-qPCR) analysis showed that ScPIP2–1 was mainly expressed in the leaf, root and bud, and its expression levels in both below- and aboveground tissues of ROC22 were up-regulated by abscisic acid (ABA), polyethylene glycol (PEG) 6000 and sodium chloride (NaCl) stresses. The chlorophyll content and ion leakage measurement suggested that ScPIP2–1 played a significant role in salt stress resistance in Nicotiana benthamiana through the transient expression test. Overexpression of ScPIP2–1 in Arabidopsis thaliana proved that this gene enhanced the salt tolerance of transgenic plants at the phenotypic (healthier state, more stable relative water content and longer root length), physiologic (more stable ion leakage, lower malondialdehyde content, higher proline content and superoxide dismutase activity) and molecular levels (higher expression levels of AtKIN2, AtP5CS1, AtP5CS2, AtDREB2, AtRD29A, AtNHX1, AtSOS1 and AtHKT1 genes and a lower expression level of the AtTRX5 gene).

Conclusions

This study revealed that the ScPIP2–1-mediated osmotic stress signaling cascade played a positive role in plant response to salt stress.

Peer Review reports

Background

Plants suffer from various environmental stresses during the process of growth and development. Salt, drought and alkali stresses are the dominant abiotic stresses in plants and can impede plant growth and development and thus production [1]. The water transport channel is one of the important defense components for resisting salt, drought and alkali stresses [2]. Regarding the water transport channel, there are three kinds of pathways: transmembrane, symplastic and apoplastic [3]. The former two accomplish water transport by cell-to-cell pathways or through plasmodesmata and membranes, and the last one transports water by water flow in the xylem [4]. However, the water transport channels could be blocked by casparian bands and suberin lamellae [3].

Aquaporins (AQPs) belong to the major intrinsic protein (MIP) family [5]. AQPs play a core role in water transport channels and are able to guide water transport across membranes between the protoplasm [6,7,8,9,10]. In plants, the molecular weights of AQPs are 23–30 kDa [5]. There are five water rings and six transmembrane α-helices in AQPs [11, 12]. According to subcellular localization and sequence similarities, AQPs in higher plants can be divided into five subfamilies: plasma membrane intrinsic proteins (PIPs), tonoplast membrane intrinsic proteins (TIPs), nodulin26-like major intrinsic proteins (NIPs), small and basic intrinsic proteins (SIPs) and X intrinsic proteins (XIPs) [13, 14]. Based on the N-terminal length of the proteins, PIPs are divided into PIP1s and PIP2s, which have different roles in plants when they respond to various stresses [5, 6, 15].

AQP genes have been cloned and characterized in several plant species, such as Arabidopsis thaliana [11], Oryza sativa [16], Zea mays [17] and Solanum lycopersicum [18]. As reported, AQPs play a major role during plant responses to water stress [6]. The expression of AQPs maintains the balance of water when plants suffer from various abiotic stresses such as salt, drought and cold [5]. PIPs are regarded as the major water transport proteins in roots and leaves of higher plants [5]. Previous research revealed that different PIP family members had various expression patterns and functions when plants were exposed to osmotic stress [19, 20]. When Arabidopsis was treated with 250 mmol·L− 1 mannitol for 2 d, the expression levels of../%25E5%258A%259E%25E5%258,525 AC/360/Youdao/Dict/8.5.2.0/resultui/html/index.html - /javascript:; AtPIP1–3, AtPIP1–4, AtPIP2–1 and AtPIP2–5 were 5 times as high as those in control leaves, but those of AtPIP1–5, AtPIP2–2, AtPIP2–3 and AtPIP2–6 decreased by 10% compared with the control [20]. Moreover, the expression levels of the PIP2 gene in Arabidopsis roots declined in response to salt stress [21]. The association between the health status of Arabidopsis and the higher expression levels of PIPs (AtPIP1–4 and AtPIP2–5) showed that the expression and activity of AtPIP1–4 and AtPIP2–5 could regulate water balance in plants after osmotic stress [22]. During the process of drought stress, the expressions of NtPIP1–1 and NtPIP2–1 were down-regulated in Nicotiana benthamiana [23]. Moderate expression levels of OsPIP1–1 can increase the seed yield, salt resistance, root hydraulic conductivity and seed germination rate of rice [24]. Furthermore, the expressions of OsPIP2 genes were positively correlated with the level of salt stress, and this phenomenon was similar to drought stress [25, 26]. In Rosa rugosa, RhPTM (PM-tethered MYB) could interact with RhPIP2–1 to maintain the water balance in cells [27]. These reports confirmed that PIPs could aid the response to salt stress in plants such as Z. mays [17], O. sativa [16] and A. thaliana [11]. However, there is no report on the identification and characterization of PIP genes in sugarcane (Saccharum spp. hybrid, var. ROC22) until now.

Sugarcane, the most important sugar crop in the world, is often planted in tropical and subtropical regions. Salt stress hinders sugarcane growth and development and also reduces its production. In the present study, first, according to the transcriptome database, eight sugarcane PIP genes (named ScPIPs) were obtained and analyzed by the bioinformatics method. Second, a full-length cDNA sequence of ScPIP2–1, homologous to the rice salt-resistant gene OsPIP2–1 (AF062393.1) [26], was cloned from the main sugarcane variety ROC22 (Saccharum spp. hybrid) in mainland China. The amino acid sequence, subcellular location and transcriptional activity of the ScPIP2–1 protein were analyzed. Third, in order to understand the expression patterns of ScPIP2–1 in ROC22, tissue-specific expression and the expression of ScPIP2–1 under polyethylene glycol (PEG) 6000, sodium chloride (NaCl) and abscisic acid (ABA) stresses were detected. Finally, the biological functions of ScPIP2–1 under salt stress were also investigated by transient overexpression of ScPIP2–1 in N. benthamiana leaves combined with stable overexpression in A. thaliana. The above results suggested that ScPIP2–1 might enhance salt tolerance in plants mainly by maintaining cell homeostasis and by eliminating reactive oxygen species (ROS). In other words, the salt tolerance of transgenic Arabidopsis could be enhanced by the ScPIP2–1-mediated osmotic stress signaling cascade.

Results

Identification and classification of sugarcane PIP family genes

Based on our previous sugarcane transcriptome, eight ScPIP family genes, namely, ScPIP1–1 (MZ362000), ScPIP1–2 (MZ362001), ScPIP1–3 (MZ362002), ScPIP1–4 (MZ362003), ScPIP2–1 (MZ362004), ScPIP2–2 (MZ362005), ScPIP2–4 (MZ362006) and ScPIP2–5 (MZ362007), were mined (Table S1). Bioinformatics analysis showed that these eight ScPIP proteins had a relative molecular mass of 30–31 kDa, isoelectric points (pIs) of 6.64–9.00, and all belonged to stable hydrophobins (Table S2). A phylogenetic tree showed that these ScPIP proteins and AQP proteins from other higher plants were divided into four subfamilies: PIPs, TIPs, NIPs and SIPs. All eight ScPIP proteins belonged to the PIP subfamily (Fig. 1).

Fig. 1
figure 1

The phylogenetic tree of sugarcane and other plant AQPs. The red filled triangle represents the proteins acquired from sugarcane in this study. The GenBank accession numbers follow the protein names. At, Arabidopsis thaliana; Sb, Sorghum bicolor; Zm, Zea mays; Os, Oryza sativa

Cloning and sequence analysis of the ScPIP2–1 gene

A full-length cDNA sequence of ScPIP2–1 was cloned from the 24-h 250 mmol·L− 1 NaCl-treated ROC22 leaves by reverse transcription-polymerase chain reaction (RT-PCR). The cDNA length of ScPIP2–1 was 1145 bp, including an open reading frame (ORF) of 873 bp that encoded 290 amino acids (Fig. 2). Moreover, the ScPIP2–1 protein had the MIP conservative structure of the NPA-motif (Asn-Pro-Ala) and the PIP feature structures (P/KDYXE/DPPP/RX3-4E/DXXELXXWSFY/WR, and GGGANXXXXGY) (Jang et al., 2004). Sequence alignment indicated that ScPIP2–1 was highly homologous with PIPs in Arabidopsis, rice, tobacco, maize and sorghum. However, the N-terminal amino acid sequences of these homologous proteins are quite different (Fig. 3).

Fig. 2
figure 2

Nucleotide acid sequences and deduced amino acid sequences of the ScPIP2–1 gene obtained by RT-PCR. The sequences of the NPA motifs are highlighted with red frames, PIP characteristic sequences are highlighted with green frames, and the primers for RT-PCR are highlighted with red underlining. *, stop codon

Fig. 3
figure 3

Amino acid sequence alignment of ScPIP2–1 and PIPs from other plant species based on DNAMAN (version 6.0.3.99, Lynnon Biosoft) software. AtPIP1–1, AY063931.1; AtPIP2–1, NM_115202.3; NtPIP1–1, AF440271.1; NtPIP2–1, NM_001326279.1; OsPIP2–1, AF062393.1; OsPIP1–1, NM_001054207; SbPIP1–2, XM_002454463.2; SbPIP2–1, XM_002461885.2; SiPIP2–1, XM_004986439.2; ZmPIP2–1, AY243801.1; ZmPIP1–1, EU967581.1. The sequences of the NPA motifs are highlighted by red rectangles

Subcellular localization analysis of the ScPIP2–1 protein

Subcellular localization analysis (Fig. 4) showed that 35S::ScPIP2–1::eGFP was expressed in the plasma membrane and cytoplasm, and 35S::eGFP was expressed in the plasma membrane, cell nucleus and cytoplasm.

Fig. 4
figure 4

Subcellular localization of ScPIP2–1 and empty vector in Nicotiana benthamiana leaves after 2 days of infiltration. Image of epidermal cell captured using green fluorescence, merged light and visible light. Red arrows mark the cell nucleus, yellow arrows mark the cytoplasm, and white arrows mark the plasma membrane

Transcription activation activity of the ScPIP2–1 protein

The DUAL membrane yeast two-hybrid system was used to detect the transcriptional activity of ScPIP2–1. Compared with the positive and negative control, pBT3-N-ScPIP2–1 did not possess transcriptional activity and showed no toxicity to the yeast strain NMY51 (Fig. 5).

Fig. 5
figure 5

Test of the ScPIP2–1 trans-activation assay. The media were added with X-Gal (5-bromo-4-chloro-3-indoxyl-D-galactopyranoside). 10− 1, 10− 2, 10− 3 and 10− 4 represent the sample concentrations diluted ten-fold, hundred-fold, thousand-fold and ten thousand-fold, respectively. SD/−Leu, the synthetic dropout medium without leucine; SD/−Leu/−Trp, the synthetic dropout medium without leucine and tryptophan; SD/−Ade/−His/−Leu/−Trp, the synthetic dropout medium without adenine, histidine, leucine and tryptophan

Tissue-specific expression of the ScPIP2–1 in different sugarcane tissues

Real time quantitative PCR (RT-qPCR) was performed to determine the expression patterns of the ScPIP2–1 gene in different sugarcane tissues, including the stem pith, epidermis, root, leaf and bud. As shown in Fig. 6, the ScPIP2–1 gene was constitutively expressed in all these sugarcane tissues but mainly in the bud, leaf and root, which was 4.71-, 2.41- and 1.55-fold higher, respectively, than that in the stem pith.

Fig. 6
figure 6

Tissue-specific expression analysis of the ScPIP2–1 gene in 10-month-old sugarcane ROC22 plants. Data were normalized to the expression level of the glyceraldehyde-3-phosphate dehydrogenase (GAPDH) gene. All data are expressed as means ± standard error (n = 3). Different lowercase letters indicate a significant difference (P < 0.05) compared to the control, as determined by Duncan’s test

Expression of the ScPIP2–1 gene in response to various abiotic stresses

In the case of ABA, NaCl and PEG stresses, except for the downregulation in aboveground tissues at 0.5 h under ABA treatment, the expression levels of the ScPIP2–1 gene in both below- and aboveground tissues of ROC22 were up-regulated (Fig. 7). During ABA stress, the ScPIP2–1 gene exhibited a sensitive response. Its expression levels peaked in belowground tissues at 6 h and in aboveground tissues at 3 h, which were 16.36- and 2.99-fold higher, respectively, than those of the control. In response to NaCl, the expression levels of the ScPIP2–1 gene were increased from 0.5 h to 3 h in belowground tissues and from 0 h to 6 h in aboveground tissues but remained unchanged at the other time points. In response to PEG 6000, the expression levels of the ScPIP2–1 gene increased at 6 h in both below- and aboveground tissues of ROC22, which were 1.59- and 2.81-fold higher, respectively, than those of the control. However, the expression levels of ScPIP2–1 remained stable at the other time points of PEG 6000 treatment. These results suggested that ScPIP2–1 had a positive response to salt and PEG 6000 stresses, most probably mediated by the ABA signaling pathway.

Fig. 7
figure 7

Relative expression of ScPIP2–1 after the application of exogenous plant hormones and abiotic stress. a The results of belowground tissues of ROC22. b The results of aboveground tissues of ROC22. ScPIP2–1 transcript abundance in ROC22 plants was assessed in the presence of 0.1 mmol·L− 1 ABA, 250 mmol·L− 1 NaCl and 25% PEG 6000. The data were normalized to the expression level of the glyceraldehyde-3-phosphate dehydrogenase (GAPDH) gene. All data are expressed as means ± standard error (n = 3). Different lowercase letters indicate a significant difference (P < 0.05), as determined with Duncan’s test. ABA, abscisic acid; NaCl, sodium chloride; PEG 6000, polyethylene glycol 6000

Transient overexpression of ScPIP2–1 gene regulated the responses of N. benthamiana leaves to salt stress

After transient overexpression of ScPIP2–1 in N. benthamiana leaves for 2 days, the overexpression tobacco plants were treated with 250 mmol·L− 1 NaCl for 24 h. Then, the salt tolerance of tobacco leaves was determined at the physiologic, phenotypic and molecular levels (Fig. 8). As shown in Fig. 8a, a darker DAB staining color was observed in the control (pCAMBIA 1301) than in ScPIP2–1 leaves. Moreover, the phenomenon of stronger salt tolerance was demonstrated by a higher SPAD index and more stable ion leakage in ScPIP2–1 overexpression plants after 24-h salt treatment (Figs. 8b, c). According to the results of Fig. 8d, the expression levels of salt-associated marker genes (NtESI3, NtDREB2 and NtP5CS) in ScPIP2–1 leaves were increased, which were 4.32-, 1.95- and 8.30-fold higher, respectively, than those of the control. These results laid the foundation for genetic transformation experiments and provided evidence for the ScPIP2–1 gene to respond to salt stress.

Fig. 8
figure 8

Effect of transient overexpression of the ScPIP2–1 gene in Nicotiana benthamiana leaves. a The chlorosis and DAB staining of N. benthamiana leaves after 24-h salt stress. b and c SPAD value and ion leakage in the N. benthamiana leaves after 24-h salt stress. SPAD represents the amount of chlorophyll, and ion leakage represents the value of conductivity. d Expression of three tobacco osmotic stress-related marker genes in the N. benthamiana leaves after 24-h salt stress. The NtEF1-α gene was used for data normalization. All data points are presented as means ± standard error (n = 3). Different lowercase letters indicate a significant difference, as determined with Duncan’s test (P < 0.05)

Overexpression of ScPIP2–1 gene improved the salt stress tolerance in transgenic Arabidopsis

Here, a functional analysis of Arabidopsis overexpression ScPIP2–1 for salt tolerance was conducted, and the indexes were measured at phenotypic, physiological and molecular levels. Three lines of homozygous transgenic Arabidopsis (OE1, OE2 and OE3) were selected in 1/2 MS medium containing hygromycin (HYG). At the phenotypic level, the transgenic Arabidopsis seedling had longer root length than wild-type Arabidopsis (WT) under 0–250 mmol·L− 1 NaCl (Fig. 9). In addition, compared to the 5-week-old transgenic Arabidopsis, the leaves of WT turned yellow and its growth was inhibited under 250 mmol·L− 1 NaCl stress (Fig. 10a). These results indicated that transgenic Arabidopsis might have better water absorption ability. As shown in Fig. 10c, the transgenic plantlets had a significantly higher relative water content (RWC) than WT plantlets after 250 mmol·L− 1 NaCl treatment.

Fig. 9
figure 9

Root length of T3 ScPIP2–1-transgenic Arabidopsis seedlings at different NaCl concentrations. a Observation of root length under different concentrations of sodium chloride (NaCl). b Measurement of root length under different concentrations of NaCl by Image J 1.8.0. c Relative water content (RWC) of Arabidopsis plantlets under treatment of 250 mmol·L− 1 NaCl. Bars with different superscripts differ significantly (P < 0.05), and error bars represent the standard error of each treatment group (n = 3). WT, wild-type Arabidopsis; OE1, OE2 and OE3, three different lines of transgenic Arabidopsis

Fig. 10
figure 10

Detection of phenotypic and physiological indexes of T3 transgenic Arabidopsis under salt stress. a The growth status of transgenic and WT plants during 250 mmmol·L− 1 sodium chloride (NaCl) treatment. b Ion leakage. c The content of malondialdehyde (MDA). d The activity of superoxide dismutase (SOD). e The content of proline (Pro). Bars with different superscripts differ significantly (P < 0.05), and error bars represent the standard error of each treatment group (n = 3). WT, wild-type Arabidopsis; OE1, OE2 and OE3, three different lines of transgenic Arabidopsis

As Figs. 10b–e and Fig. 11 show, the transgenic Arabidopsis exhibited a stronger salt tolerance at the physiological and molecular levels. It was noted that the transgenic Arabidopsis had a higher proline (Pro) content, and the expression levels of Pro synthesis-related genes AtP5CS1 and AtP5CS2 were up-regulated by salt stress (Fig. 10e and Figs. 11c, d). According to the results of Fig. 10b and Figs. 11a, e and f, the transgenic Arabidopsis had a more stable intracellular environment than WT, which was exhibited by the insignificantly increased ion leakage (3–7 days) and the higher expression levels of osmotic stress-defense genes AtKIN2, AtDREB2 and AtRD29A. Three genes, AtNHX1, AtSOS1 and AtHKT1, which correlated with ion transport, were all up-regulated after 250 mmol·L− 1 NaCl treatment (Figs. 11g, h, i). These phenomena confirmed that overexpression of ScPIP2–1 could enhance ion exchange to improve the salt tolerance of transgenic plants. Regarding the correlation between ROS removal and salt tolerance, the transgenic Arabidopsis had less malondialdehyde (MDA) and stronger superoxide dismutase (SOD) activity than WT, and the expression levels of the AtTRX5 gene, which is related to ROS removal, were down-regulated, suggesting that overexpression of ScPIP2–1 could restrict AtTRX5 expression by taking part in ROS elimination (Figs. 10c, d and Fig. 11b).

Fig. 11
figure 11

Expression of salt stress-related genes in ScPIP2–1-transgenic Arabidopsis based on RT-qPCR. AtKIN2, AtDREB2 and AtRD29A represent the osmotic stress-defense genes. AtTRX5 represents the reactive oxygen species (ROS) elimination gene. AtP5CS1 and AtP5CS2 represent the proline (Pro) synthesis-related genes. AtNHX1, AtSOS1 and AtHKT1 represent the ion exchange-correlated genes. Arabidopsis was grown in soil and was treated with 250 mmol·L− 1 sodium chloride (NaCl). Bars with different superscripts differ significantly (P < 0.05), and error bars represent the standard error of each treatment group (n = 3). WT, wild-type Arabidopsis; OE1, OE2 and OE3, three different lines of transgenic Arabidopsis

Discussion

AQPs play important roles as water channels in regulating water balance [6,7,8,9,10]. At present, many AQPs from various plant species have been used in transgenic research to improve salt stress tolerance in plants [28,29,30,31]. In this study, eight AQPs, all of which belong to the PIP subfamily, were mined from the sugarcane transcriptome database (Fig. 1). Johanson et al. suggested that the naming of AQPs should refer to the protein subcellular localization and sequence homology [11]. In subsequent research, the AQP proteins located in the plasma membrane and cytoplasm were renamed as PIPs in higher plants [5, 11]. Our research revealed that ScPIP2–1 was localized in the plasma membrane and cytoplasm (Fig. 4), which is in accordance with previous studies [27, 29]. Protein-protein interactions play an important role in plant development and growth [32]. Based on previous research in roses, RhPTM could interact with RhPIP2–1, indicating its important role in maintaining water balance [27]. In the present study, pBT3-N-ScPIP2–1 did not possess transcriptional activity and showed no toxicity to the yeast strain NMY51 (Fig. 5), suggesting that it could be used for yeast two-hybrid screening in future research.

Temporal and spatial expression of AQPs often represents their diverse functions [9]. SbAQPs were mainly expressed in roots and played major roles in sorghum under waterlogging [33]. AtPIP1–4 and AtPIP2–5 were mainly expressed in the leaf and root when Arabidopsis was exposed to salt and drought stresses [22]. Likewise, the tissue-specific expression of ScPIP2–1 (Fig. 6) proved that it may play an important role in maintaining water transport in leaf and root tissues. Accordingly, we speculated that ScPIP2–1 also participated in the sugarcane cell development due to the fact that ScPIP2–1 was mainly expressed in bud tissues of sugarcane (Fig. 6), which is similar to the phenomenon that RhPIP2–1 participated in rose petal cell swelling with higher expression in the meristem [34]. ABA is a major phytohormone that can enhance the tolerance of osmotic stress in plants [35]. In this study, the expression levels of the ScPIP2–1 gene in both below- and aboveground tissues of ROC22 were increased by ABA treatment (Fig. 7). The transcripts of the ScPIP2–1 gene were also up-regulated under NaCl and PEG 6000 stresses (Fig. 7). The above phenomena were the same as in many previous studies, for example, the expression levels of ClPIP were up-regulated under ABA treatment and had a positive response to osmotic stresses in Citrullus lanatus [36], and PvPIP2–9 withstood salt and drought stresses through ABA signal regulation in Panicum virgatum [37]. Water deficit can cause difficulty in root water absorption, mainly due to a lower osmotic potential. The research demonstrates that ABA can regulate stomatal closing to reduce the rate of water loss [38]. Taking all these findings into consideration, we believe that ScPIP2–1 may have a positive response to PEG 6000 and NaCl stresses through the ABA signal in sugarcane.

What should be stressed here is that we found the overexpression of ScPIP2–1 in Arabidopsis improved its tolerance to salt stress, which was correlated with the higher expression levels of salt-associated marker genes (Fig. 11), lower levels of stress-induced ROS (Fig. 10c), higher activity of ROS scavenging enzyme (Fig. 10d) and higher expression levels of Pro biosynthetic genes and a higher Pro content (Fig. 10e and Figs. 11c, d). In particular, under salt stress, ScPIP2–1 overexpression Arabidopsis displayed healthier physiological states (Fig. 10a), longer length of roots (Fig. 9) and more stable cell ion leakage (Fig. 10b) compared with WT, which suggested that the transgenic Arabidopsis had more stable cell pressure. Although the ScPIP2–1 overexpression improved salt tolerance in Arabidopsis in our study, AtPIP2–1 overexpression resulted in reduced salt stress tolerance in Arabidopsis and tobacco [39]. These seemingly conflicting results are not uncommon. In fact, many AQP overexpression studies have produced contrasting results, for example, AQP overexpression plants showed either positive or negative effects due to stress tolerance [9, 40]. There are differences in gene regulation, even among highly homologous AQPs, indicating their divergent functions [39]. Another explanation for the contrasting results might lie in the differences in protein sequences between ScPIP2–1 and other plant PIP proteins. Amino acid sequence differences were mainly found in the N-terminus, which was expected to be exposed on the cytosol side [41]. From all the above, it was tempting to infer that contrasting stress phenotypes may be due to differences in the N-termini of the PIP protein (Fig. 3), which contains important motifs such as those for activity regulation, protein stability, protein interaction or even subcellular localization.

In the salt tolerance analysis of transgenic Arabidopsis, we found up-regulation of AtP5CS1 and AtP5CS2 (Figs. 11c, d), higher Pro content (Fig. 10e) and increased activity of SOD (Fig. 10d) with reduced levels of MDA (Fig. 10c) in ScPIP2–1 overexpression Arabidopsis under salt stress. Pro mainly functions in defense and turgor pressure maintenance against water-deprived conditions [42]. The expression levels of three ion exchange-correlated genes (AtNHX1, AtSOS1 and AtHKT1) were up-regulated, which may improve the salt tolerance of ScPIP2–1 overexpression plants, which is in accordance with a previous report that AtNHX1, AtSOS1 and AtHKT1 played important roles in salt stress resistance [43]. Hence, the enhanced salt stress tolerance in ScPIP2–1 overexpression Arabidopsis seemed to result in, at least in part, the increased expression of Pro biosynthetic genes and elevated activities of ROS scavenging enzymes, as well as the down-regulation of AtTRX5 (Fig. 11b). Similar to our findings, the enhanced stress tolerance in MsPIP2–2 overexpression transgenic Arabidopsis was correlated with decreased levels of MDA and increased Pro contents and activity of the SOD enzyme [29]. In the present study, salt tolerance improvement by up-regulation of salt stress-regulated genes and an increase in ROS scavenging enzyme activity suggested that ScPIP2–1 overexpression might sensitize transgenic Arabidopsis, causing the overexpression Arabidopsis react faster to salt stress signals and eventually enhance stress defense. According to the results of Figs. 11e and f, which indicated the up-regulation of AtDREB2 and AtRD29A genes under salt stress, we speculate that ScPIP2–1 in transgenic Arabidopsis could activate salt stress sensing or the upstream steps in signaling pathways to induce a better stress-tolerance mechanism than in WT. Consistent with our speculation, it has been suggested that AQPs may be part of an osmotic stress signaling cascade [9]. Additionally, it has even been proposed that AQPs may act as osmosensors [44]. Further study will be required to investigate this possibility.

Conclusions

In the present study, eight ScPIP genes that belong to the PIP subfamily of AQPs were mined from our previous sugarcane transcriptome database. A full-length cDNA of ScPIP2–1 was then cloned from ROC22. ScPIP2–1 was localized in the plasma membrane and cytoplasm. The ScPIP2–1 gene was constitutively expressed in sugarcane tissues, and its transcript levels were increased by ABA, NaCl and PEG 6000 treatments. Yeast two-hybrid screening demonstrated that pBT3-N-ScPIP2–1 did not possess transcriptional activity and had no toxicity to the yeast strain NMY51. The transient overexpression of ScPIP2–1 in N. benthamiana leaves could enhance the salt tolerance. Moreover, overexpression ScPIP2–1 in Arabidopsis could enhance the salt tolerance of transgenic plants at the phenotypic, physiological and molecular levels. These results would be helpful in mining and functional identification of ScPIP family genes in sugarcane in future.

Methods

Plant materials and treatments

One panel of sugarcane material was used in this study. The main sugarcane variety ROC22 (Saccharum spp. hybrid) in mainland China was provided by the Key Laboratory of Sugarcane Biology and Genetic Breeding, Ministry of Agriculture and Rural Affairs (Fuzhou, China). For tissue-specific expression analysis, 10-month-old ROC22 plants were selected. The samples included the root, bud, + 1 leaf (the youngest fully expanded leaf with a visible dewlap), stem pith and epidermis [45]. For osmotic stress and exogenous hormone treatments, 4-month-old ROC22 plants were cultivated hydroponically for 7 days. The cultivation conditions were 28 °C with 16-h light and 8-h darkness. Then, 25 PEG 6000, 250 mmol·L− 1 NaCl and 0.1 mmol·L− 1 ABA were used as simulated environmental stress conditions [46,47,48]. The leaf and root of each treatment were collected at 0, 0.5, 3, 6 and 24 h. Three biological replicates of each sample were prepared, and all the above-mentioned samples were fixed in liquid nitrogen and stored at − 80 °C until total RNA extraction.

Total RNA extraction and the first strand of cDNA synthesis

Total RNA was extracted using TRIeasy™ Total RNA Extraction Reagent (YEASEN, Shanghai, China), and 1.0 μg total RNA was used to synthesize first-strand cDNA using the Hifair® 1st Strand cDNA Synthesis Kit (gDNA digester plus) (YEASEN, Shanghai, China). The cDNA was detected by 1% gel electrophoresis [49]. The cDNA of 24-h 250 mmol·L− 1 NaCl-treated ROC22 leaves was used for gene cloning.

Identification and bioinformatics analysis of ScPIP family genes

Eight ScPIP genes were minded from our previous sugarcane transcriptome database [50, 51]. The sequences of these genes were translated and analyzed using the Open Reading Frame Finder (https://www.ncbi.nlm.nih.gov/orffinder/) and DNAMAN V6.0.3.99 software. The signal peptides and pIs were predicted using ExPASy (http://us.expasy.org/tools). A phylogenetic tree was constructed with the maximum likelihood (ML) method (1000 bootstrap replicates) using the MEGA X 10.0.5 program [16, 52].

The full-length cDNA sequence of ScPIP2–1, which was minded above, was cloned by RT-PCR from 24-h 250 mmol·L− 1 NaCl-treated leaf tissue of ROC22. The amplification procedure was set as 94 °C for 5 min, followed by 35 cycles of 94 °C for 30 s, 52 °C for 30 s and 72 °C for 1 min, and then 72 °C for 10 min in an elongation step. The PCR products were purified by a gel-purification kit (Magen, Guangzhou, China), cloned into the pMD 19-T vector (Takara, Dalian, China) and sent for sequencing.

Subcellular localization assay

The ORF of ScPIP2–1 without the stop codon was inserted between two restriction sites (BamH I and Xba I) of the 35S::eGFP vector (pCAMBIA 2300-eGFP). Then, the recombinant vector of 35S::ScPIP2–1::eGFP was introduced into Agrobacterium GV3101. Agrobacterium-mediated transient expression was performed in N. benthamiana leaves [53,54,55]. Subcellular localization of the GFP reporter protein was determined by confocal laser scanning microscopy using a Leica TCS SP5 (Wetzlar, Germany).

Self-activation of the ScPIP2–1 in yeast

To analyze the transcriptional activity of ScPIP2–1, the DUAL membrane yeast two-hybrid system (including three reporter genes, HIS3, ADE2 and lacZ) was used following the manufacturer’s directions for the DUAL yeast two-hybrid system and according to the protocol established in our group [48]. The ScPIP2–1 gene was PCR-amplified from pMD19-T-ScPIP2–1 using primers ScPIP2–1 pBT3-F and ScPIP2–1 pBT3-R (Table S3). Then, Sfi I endonuclease was used to construct the bait vector pBT3-N-ScPIP2–1. NMY51 strains carrying the treatment (positive plasmid pBT3-N-ScPIP2–1), the negative control (empty vector pBT3-N and empty vector pRR3-N) and the positive control (pNubG-Fe65 and pTSU2-App) vectors were cultured in synthetic dropout medium without leucine and tryptophan (SD/−Leu/−Trp) for 2 days at 30 °C at 150 rpm [48].

In order to verify the self-activation activity of NMY51 strains plus positive plasmid pBT3-N-ScPIP2–1, 8.0 μL yeasts plus 1.0 μL X-gal were cultured on solid media, which were SD/−Leu (the synthetic dropout medium without leucine), SD/−Leu/−Trp and SD/−Ade/−His/−Leu/−Trp (the synthetic dropout medium without adenine, histidine, leucine and tryptophan). The sample was diluted tenfold (10− 1), one hundred fold (10− 2), one thousand fold (10− 3) and ten thousand fold (10− 4).

Expression of the ScPIP2–1 gene in different sugarcane tissues and in response to various abiotic stresses

RT-qPCR analysis was conducted to reveal the expression patterns of ScPIP2–1 in sugarcane tissues (root, bud, leaf, stem pith and stem epidermis) and in response to various abiotic stresses (NaCl, PEG 6000 and ABA). Glyceraldehyde-3-phosphate dehydrogenase (GAPDH, GenBank Accession Number: CA254672) was used as a reference gene (Table S3) [56]. The specific RT-qPCR primers of ScPIP2–1 (Table S3) were designed using Beacon Designer 8.13 software. An ABI 7500 real-time PCR system (Applied Biosystems, Foster City, CA, USA) and Hieff™ qPCR SYBR Green Master Mix (Low Rox) (YEASEN, Shanghai, China) were used for RT-qPCR analysis. The RT-qPCR reaction system was subjected to 50 °C for 2 min, 95 °C for 10 min, 95 °C for 15 s and 59 °C for 1 min, for 40 cycles. A melting curve analysis was performed at 95 °C for 15 s, 60 °C for 1 min, 95 °C for 15 s and 60 °C for 30 s. Three biological and three technical replicates were performed. The standard curve plot and amplification efficiency of ScPIP2–1 and GAPDH genes in RT-qPCR analysis were showed in Fig. S1 and Table S4. The 2−∆∆CT method [54], SPSS 9 software and Origin 8 software were used to calculate the relative expression of the target gene, to analyze the significance level of the experimental data and to structure the histogram.

Transient expression of ScPIP2–1 in N. benthamiana

To investigate the function of ScPIP2–1 under high salt treatment in N. benthamiana, the overexpression vector pCAMBIA 1301-ScPIP2–1 was constructed using T4 ligase (YEASEN, Shanghai, China). The recombinant vector of pCAMBIA 1301-ScPIP2–1 and the empty vector of pCAMBIA 1301 were separately introduced into Agrobacterium GV3101. Agrobacterium-mediated transient expression in the N. benthamiana leaf was performed [52,53,54]. The six-leaf stage N. benthamiana plants were irrigated with 50 mL of 250 mmol·L− 1 NaCl every day. All materials were cultivated at 24 °C with an illumination intensity of 15,000 μmol·m− 2·s− 1, a photoperiod of 16-h light and 8-h darkness and a relative humidity of 65%. The histocytological analysis was processed by 3, 3′-diaminobenzidine (DAB) staining at 0 h and 24 h on leaves under NaCl stress. At the same time, the chlorophylls and ion leakage of the leaves were determined using soil and plant analyzer development measurements (SPAD, KONICA MINOLTA, Japan) and the soaking method [57], respectively. The expression levels of salt-associated genes, i.e., NtESI3 (salt stress-induced hydrophobic peptide gene) [58], NtDREB2 (DREB transcription factor) [59] and NtP5CS (correlation Pro production gene) [60] (Table S3), were detected by RT-qPCR. The NtEF1-α gene (Table S3) was used as an internal reference for normalization. The standard curve plot and amplification efficiency of RT-qPCR primers were showed in Fig. S1 and Table S4. The 2−∆∆CT method [54], SPSS 9 software and Origin 8 software were used to calculate the relative expression of the target gene, to analyze the significance level of the experimental data and to structure the histogram.

Stable expression of ScPIP2–1 in Arabidopsis and the salt tolerance analysis of transgenic plants

The ScPIP2–1 gene was genetically transformed into Arabidopsis by Agrobacterium-mediated transformation. The overexpression vector was pCAMBIA 1301, which contained a marked gene for HYG resistance. Seeds of transgenic Arabidopsis lines of T1, T2 and T3 generations were germinated and selected on 1/2 MS medium with 25 mg·L− 1 HYG. As a control, WT seeds were also placed on 1/2 MS medium with HYG. Two weeks later, the seedlings were transferred to soil pots and cultured at 22 °C with an illumination intensity of 15,000 μmol·m− 2·s− 1, a photoperiod of 16-h light and 8-h darkness and a relative humidity of 65%. The seeds of T1, T2 and T3 generations were generated and collected.

To determine the root lengths, the seeds of WT and three lines of T3 transgenic Arabidopsis (OE1, OE2 and OE3) were separately placed on 1/2 MS medium containing 0, 50, 150, 200 or 250 mmol·L− 1 NaCl. A Nikon camera was used to take the photos 7 days later, and Image J 1.8.0 was used to measure the root lengths [47, 61].

In order to analyze the salt tolerance of ScPIP2–1 overexpression Arabidopsis at physiological and molecular levels, the 5-week plants of WT and three lines of T3 transgenic Arabidopsis (OE1, OE2 and OE3) were cultivated in peat medium and treated with 50 mL of 250 mmol·L− 1 NaCl every day. The Nikon camera was used to take photos at 0, 3 and 7 days post NaCl treatment. The leaf samples for the analysis of physiological indexes were collected at 0 d and 7 d post treatment, and other leaf samples for the analysis of molecular indexes were collected at 0 d, 3 d and 7 d post treatment. Three biological replicates of each sample were prepared. Three plantlets, randomly selected, were defined as one biological replicate, and three biological replicates of each sample were prepared to measure the fresh weight (FW) and dry weight (DW). The RWC was calculated by RWC = (FW-DW) × FW− 1 [62]. The contents of MDA and Pro as well as the enzymatic activity of SOD were determined by spectrophotometry using MDA, Pro and SOD assay kits (Solarbio, Beijing, China) [47, 61]. The ion leakage was determined by the soaking method [57]. The expressions of the osmotic stress-defense genes (AtKIN2, AtDREB2 and AtRD29A), the ROS-elimination gene (AtTRX5), the Pro synthesis-related genes (AtP5CS1 and AtP5CS2) and the ion exchange-related genes (AtNHX1, AtSOS1 and AtHKT1) in Arabidopsis (Table S3) were analyzed by RT-qPCR [29]. The AtActin gene (Table S3) was used as an internal reference for normalization [29]. The standard curve plot and amplification efficiency of these RT-qPCR primers were showed in Fig. S1 and Table S4. The 2−∆∆CT method [54], SPSS 9 software and Origin 8 software were used to calculate the relative expression of the target gene, to analyze the significance level of the experimental data and to structure the histogram.

Availability of data and materials

The datasets supporting the conclusions of this article are available in the NCBI repository (https://www.ncbi.nlm.nih.gov/) with the GenBank accession numbers of MZ362000, MZ362001, MZ362002, MZ362003, MZ362004, MZ362005, MZ362006 and MZ362007. This is also to confirm that we have got the administrative permission to access and use the sugarcane transcriptome database reported by Zhou et al. [50] and Yang et al. [51].

Abbreviations

AQPs:

Aquaporins

MIP:

Major intrinsic protein

PIPs:

Plasma membrane intrinsic proteins

TIPs:

Tonoplast membrane intrinsic proteins

NIPs:

Nodulin26-like major intrinsic proteins

SIPs:

Small and basic intrinsic proteins

XIPs:

X intrinsic proteins

PEG:

Polyethylene glycol

NaCl:

Sodium chloride

ABA:

Abscisic acid

ROS:

Reactive oxygen species

RT-PCR:

Reverse transcription-polymerase chain reaction

ORF:

Open reading frame

pI:

Isoelectric point

RT-qPCR:

Real time quantitative PCR

RWC:

Relative water content

Pro:

Proline

MDA:

Malondialdehyde

SOD:

Superoxide dismutase

HYG:

Hygromycin

FW:

Fresh weight

DW:

Dry weight

GAPDH:

Glyceraldehyde-3-phosphate dehydrogenase

References

  1. Boyer JS. Plant productivity and environment. Science. 1982;218:443–8.

    Article  CAS  PubMed  Google Scholar 

  2. Aharon R, Shahak Y, Wininger S, Bendov R, Kapulnik Y, Galili G. Overexpression of a plasma membrane aquaporin in transgenic tobacco improves plant vigor under favorable growth conditions but not under drought or salt stress. Plant Cell. 2003;15:439–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Steudle E. Water uptake by roots: effects of water deficit. J Exp Bot. 2000;51:1531–42.

    Article  CAS  PubMed  Google Scholar 

  4. Steudle E, Ranathunge K. Apoplastic water transport in roots. Springer Neth. 2007;9:119–30.

    Google Scholar 

  5. Maurel C, Boursiac Y, Luu D, Santoni V, Shahzad Z, Verdoucq L. Aquaporins in plants. Physiol Rev. 2015;95:1321–58.

    Article  CAS  PubMed  Google Scholar 

  6. Chaumont F, Tyerman SD. Aquaporins: highly regulated channels controlling plant water relations. Plant Physiol. 2014;164:1600–18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Kaldenhoff R, Ribascarbo M, Sans JF, Lovisolo C, Heckwolf M, Uehlein N. Aquaporins and plant water balance. Plant Cell Environ. 2008;31:658–66.

    Article  CAS  PubMed  Google Scholar 

  8. Knepper MA. The aquaporin family of molecular water channels. Proc Natl Acad Sci U S A. 1994;91:6255–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Maurel C. Plant aquaporins: novel functions and regulation properties. FEBS Lett. 2007;581:2227–36.

    Article  CAS  PubMed  Google Scholar 

  10. Tyerman SD, Niemietz CM, Bramley H. Plant aquaporins: multifunctional water and solute channels with expanding roles. Plant Cell Environ. 2002;25:173–94.

    Article  CAS  PubMed  Google Scholar 

  11. Johanson U, Karlsson M, Johansson I, Gustavsson S, Kjellbom PJ. The complete set of genes encoding major intrinsic proteins in Arabidopsis provides a framework for a new nomenclature for major intrinsic proteins in plants. Plant Physiol. 2001;126:1358–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Susanna TH, Yi W, Kristina H, Urban J, Maria K, Emad T, et al. Structural mechanism of plant aquaporin gating. Nature. 2006;439:688–94.

    Article  Google Scholar 

  13. Quigley F, Rosenberg JM, Shacharhill Y, Bohnert HJ. From genome to function: the Arabidopsis aquaporins. Genome Biol. 2001;3:1–17.

    Article  Google Scholar 

  14. Wallace IS, Choi W, Roberts DM. The structure, function and regulation of the nodulin 26-like intrinsic protein family of plant aquaglyceroporins. Plant Cell Environ. 2006;1758:1165–75.

    CAS  Google Scholar 

  15. Hachez C, Besserer A, Chevalier AS, Chaumont F. Insights into plant plasma membrane aquaporin trafficking. Trends Plant Sci. 2013;18:344–52.

    Article  CAS  PubMed  Google Scholar 

  16. Sakurai J, Ishikawa F, Yamaguchi T, Uemura M, Maeshima M. Identification of 33 rice aquaporin genes and analysis of their expression and function. Plant Cell Physiol. 2005;46:1568–77.

    Article  CAS  PubMed  Google Scholar 

  17. Bari A, Farooq M, Hussain A, Qamar T, Muhammad, Abbas MW, et al. Genome-wide bioinformatics analysis of aquaporin gene family in maize (Zea mays L.). J Phylogenet. Evol. Biol. 2018;6:2.

  18. Reuscher S, Akiyama M, Mori C, Aoki K, Shibata D, Shiratake K. Genome-wide identification and expression analysis of aquaporins in tomato. PLoS One. 2013;8:e79052.

    Article  PubMed  PubMed Central  Google Scholar 

  19. Galmes J, Pou A, Alsina MM, Tomas M, Medrano H, Flexas J. Aquaporin expression in response to different water stress intensities and recovery in Richter-110 (Vitis sp.): relationship with ecophysiological status. Planta. 2007;226:671–81.

    Article  CAS  PubMed  Google Scholar 

  20. Jang JY, Kim DG, Kim YO, Kim JS, Kang H. An expression analysis of a gene family encoding plasma membrane aquaporins in response to abiotic stresses in Arabidopsis thaliana. Plant Mol Biol. 2004;54:713–25.

    Article  CAS  PubMed  Google Scholar 

  21. Boursiac Y, Chen S, Luu D, Sorieul M, Dries NVD, Maurel C. Early effects of salinity on water transport in Arabidopsis roots. Molecular and cellular features of aquaporin expression. Plant Physiol. 2005;139:790–805.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Alexandersson E, Fraysse L, Sjövall-Larsen S, Gustavsson S, Fellert M, Karlsson M, et al. Whole gene family expression and drought stress regulation of aquaporins. Plant Mol Biol. 2005;59:469–84.

    Article  CAS  PubMed  Google Scholar 

  23. Mahdieh M, Mostajeran A, Horie T, Katsuhara M. Drought stress alters water relations and expression of PIP-type aquaporin genes in Nicotiana tabacum plants. Plant Cell Physiol. 2008;49:801–13.

    Article  CAS  PubMed  Google Scholar 

  24. Liu C, Fukumoto T, Matsumoto T, Gena P, Frascaria D, Kaneko T, et al. Aquaporin OsPIP1;1 promotes rice salt resistance and seed germination. Plant Physiol Biochem. 2013;63:151–8.

    Article  CAS  PubMed  Google Scholar 

  25. Ding L, Uehlein N, Kaldenhoff R, Guo S, Zhu Y, Kai L. Aquaporin PIP2;1 affects water transport and root growth in rice (Oryza sativa L.). Plant Physiol Biochem. 2019;139:152–60.

    Article  CAS  PubMed  Google Scholar 

  26. Guo L, Wang Z, Lin H, Cui W, Chen J, Liu M, et al. Expression and functional analysis of the rice plasma-membrane intrinsic protein gene family. Cell Res. 2006;16:277–86.

    Article  CAS  PubMed  Google Scholar 

  27. Zhang S, Feng M, Chen W, Zhou X, Lu J, Wang Y, et al. In rose, transcription factor PTM balances growth and drought survival via PIP2;1 aquaporin. Nat Plants. 2019;5:290–9.

    Article  CAS  PubMed  Google Scholar 

  28. Alavilli H, Awasthi JP, Rout GR, Sahoo L, Lee B, Panda SK. Overexpression of a barley aquaporin gene, HvPIP2;5 confers salt and osmotic stress tolerance in yeast and plants. Front Plant Sci. 2016;7:1566.

    Article  PubMed  PubMed Central  Google Scholar 

  29. Li S, Liu J, An Y, Gao Y, Yang P. MsPIP2;2, a novel aquaporin gene from Medicago sativa, confers salt tolerance in transgenic Arabidopsis. Environ Exp Bot. 2019;165:39–52.

    Article  CAS  Google Scholar 

  30. Wang X, Gao F, Bing J, Sun W, Feng X, Ma X, et al. Overexpression of the jojoba aquaporin gene, ScPIP1, enhances drought and salt tolerance in transgenic Arabidopsis. Int J Mol Sci. 2019;20(1):153.

    Article  PubMed Central  Google Scholar 

  31. Zhou L, Zhou J, Xiong Y, Liu C, Wang J, Wang G, et al. Overexpression of a maize plasma membrane intrinsic protein ZmPIP1;1 confers drought and salt tolerance in Arabidopsis. PLoS One. 2018;13:e198639.

    Google Scholar 

  32. Brückner A, Polge C, Lentze N, Auerbach D, Schlattner U. Yeast two-hybrid, a powerful tool for systems biology. Int J Mol Sci. 2009;10:2763–88.

    Article  PubMed  PubMed Central  Google Scholar 

  33. Kadam S, Abril A, Dhanapal AP, Koester RP, Vermerris W, Jose S, et al. Characterization and regulation of aquaporin genes of sorghum [sorghum bicolor (L.) Moench] in response to waterlogging stress. Front. Plant Sci. 2017;8:862.

    Google Scholar 

  34. Ma N, Xue J, Li Y, Liu X, Dai F, Jia W, et al. Rh-PIP2;1, a rose aquaporin gene, is involved in ethylene-regulated petal expansion. Plant Physiol. 2008;148:894–907.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Song J, Ye G, Qian Z, Ye Q. Virus-induced plasma membrane aquaporin PsPIP2;1 silencing inhibits plant water transport of Pisum sativum. Bot Stud. 2016;57:15.

    Article  PubMed  PubMed Central  Google Scholar 

  36. Zhou Y, Tao J, Ahammed GJ, Li J, Yang Y. Genome-wide identification and expression analysis of aquaporin gene family related to abiotic stress in watermelon. Genome. 2019;62:643–56.

    Article  CAS  PubMed  Google Scholar 

  37. Zhang J, Wen W, Li H, Lu Q, Xu B, Huang B. Overexpression of an aquaporin gene PvPIP2;9 improved biomass yield, protein content, drought tolerance and water use efficiency in switchgrass (Panicum virgatum L.). GCB Bioenergy. 2020;12:979–91.

    Article  CAS  Google Scholar 

  38. Zhang F, Frances S, Nichols DS, Cardoso AA, Brodribb TJ, Mcadm SAM. Leaves, not roots or floral tissue, are the main site of rapid, external pressure-induced ABA biosynthesis in angiosperms. J Exp Bot. 2018;5:1261–7.

    Article  Google Scholar 

  39. Ji Y, Seong HL, Ji Y, Gap CC, Sung JA, Hunseun K. Transgenic Arabidopsis and tobacco plants overexpressing an aquaporin respond differently to various abiotic stresses. Plant Mol Biol. 2007;64:621–32.

    Article  Google Scholar 

  40. Zhou L, Wang C, Liu R, Han Q, Vandeleur RK, Du J, et al. Constitutive overexpression of soybean plasma membrane intrinsic protein GmPIP1;6 confers salt tolerance. BMC Plant Biol. 2014;14:181.

    Article  PubMed  PubMed Central  Google Scholar 

  41. Walz T, Hirai T, Murata K, Heymann JB, Mitsuoka K, Fujiyoshi Y, et al. The three-dimensional structure of aquaporin-1. Nature. 1997;387:624–7.

    Article  CAS  PubMed  Google Scholar 

  42. Ashraf M, Foolad MR. Roles of glycine betaine and proline in improving plant abiotic stress resistance. Environ Exp Bot. 2007;59:206–16.

    Article  CAS  Google Scholar 

  43. Zhu J. Regulation of ion homeostasis under salt stress. Curr Opin Plant Biol. 2003;6:441–5.

    Article  CAS  PubMed  Google Scholar 

  44. Hill AB, Shacharhill B, Shacharhill Y. What are aquaporins for? J Membr Biol. 2001;197:1–32.

    Article  Google Scholar 

  45. Sun T, Liu F, Wang W, Wang L, Wang Z, Li J, et al. The role of sugarcane catalase gene ScCAT2 in the defense response to pathogen challenge and adversity stress. Int J Mol Sci. 2018;19:2686.

    Article  PubMed Central  Google Scholar 

  46. Su Y, Guo J, Ling H, Chen S, Wang S, Xu L, et al. Isolation of a novel peroxisomal catalase gene from sugarcane, which is responsive to biotic and abiotic stresses. PLoS One. 2014;9:e84426.

    Article  PubMed  PubMed Central  Google Scholar 

  47. Su Y, Wang Z, Liu F, Li Z, Peng Q, Guo J, et al. Isolation and characterization of ScGluD2, a new sugarcane beta-1,3-glucanase D family gene induced by Sporisorium scitamineum, ABA, H2O2, NaCl, and CdCl2 stresses. Front Plant Sci. 2016;7:1348.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Su W, Ren Y, Wang D, Su Y, Feng J, Zhang C, et al. The alcohol dehydrogenase gene family in sugarcane and its involvement in cold stress regulation. BMC Genomics. 2020;21:521.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Wang L, Liu F, Zhang X, Wang W, Sun T, Chen Y, et al. Expression characteristics and functional analysis of the ScWRKY3 gene from sugarcane. Int J Mol Sci. 2018;19:4059.

    Article  PubMed Central  Google Scholar 

  50. Zhou D, Liu X, Gao S, Guo J, Su Y, Ling H, et al. Foreign cry1Ac gene integration and endogenous borer stress-related genes synergistically improve insect resistance in sugarcane. BMC Plant Biol. 2018;18:342.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Yang Y, Gao S, Su Y, Lin Z, Guo J, Li M, et al. Transcripts and low nitrogen tolerance regulatory and metabolic pathways in sugarcane under low nitrogen stress. Environ Exp Bot. 2019;163:97–111.

    Article  CAS  Google Scholar 

  52. Mao H, Wang W, Su W, Su Y, Liu F, Li C, et al. Genome-wide identification, phylogeny, and expression analysis of Sec14-like PITP gene family in sugarcane. Plant Cell Rep. 2019;38:637–55.

    Article  Google Scholar 

  53. Su Y, Xu L, Wang S, Wang Z, Yang Y, Chen Y, et al. Identification, phylogeny, and transcript of chitinase family genes in sugarcane. Sci Rep. 2015;5:–10708.

  54. Livak KJ, Schmittgen TD. Analysis of relative gene expression data using realtime quantitative PCR and the 2-∆∆CT method. Methods. 2001;25:402–8.

    Article  CAS  PubMed  Google Scholar 

  55. Que Y, Xu L, Xu J, Zhang J, Zhang M. Selection of control genes in real-time qPCR analysis of gene expression in sugarcane. Chin J Trop Crops. 2009;30:274–8.

    Google Scholar 

  56. Sreedharan S, Shekhawat UKS, Ganapathi TR. Constitutive and stress-inducible overexpression of a native aquaporin gene (MusaPIP2;6) in transgenic banana plants signals its pivotal role in salt tolerance. Plant Mol Biol. 2015;88:41–52.

    Article  CAS  PubMed  Google Scholar 

  57. Brunetti SC, Arseneault MKM, Gulick PJ. Characterization of the ESI3/RCI2/PMP3 gene family in the Triticeae. BMC Genomics. 2018;19:898.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Nakashima K, Shinwari ZK, Sakuma Y, Seki M, Miura S, Shinozaki K, et al. Organization and expression of two Arabidopsis DREB2 genes encoding DRE-binding proteins involved in dehydration- and high-salinity-responsive gene expression. Plant Mol Biol. 2000;42:657–65.

    Article  CAS  PubMed  Google Scholar 

  59. Razavizadeh R, Ehsanpour AA, Ahsan N, Komatsu S. Proteome analysis of tobacco leaves under salt stress. Peptides. 2009;30:1651–9.

    Article  CAS  PubMed  Google Scholar 

  60. Ren Y, Mao H, Su W, Su Y, Zhang C, Shi M, et al. Improved drought and salt tolerance in transgenic Nicotiana benthamiana by overexpressing sugarcane ScSEC14p gene. Sugar Tech. 2020;1-10.

  61. Smart RE, Bingham GE. Rapid estimates of relative water content. Plant Physiol. 1974;53:258–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgments

We are grateful to the reviewers for their helpful comments on the original manuscript. We would like to thank editors for their efficient works.

Funding

This research was funded by the National Key R&D Program of China (2019YFD1000500), the Opening Foundation of Key Laboratory of Pathogenic Fungi and Mycotoxins of Fujian Province (144/PTJH13004) and the China Agriculture Research System (CARS-17). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

Conceived and designed the experiments: YS and YQ conceived and designed the research. HT and ZL prepared the materials. HT, QY, FL, WS, HL, JL and YS conducted the experiments. HT, QY and CZ analyzed the data. HT wrote the manuscript. YS and YQ revised and approved the manuscript. The authors read and approved the final manuscript.

Corresponding authors

Correspondence to Yachun Su or Youxiong Que.

Ethics declarations

Ethics approval and consent to participate

Not applicable. This is to confirm that no specific permits were needed for the described experiments, and this study did not involve any endangered or protected species.

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: Table S1.

The nucleic acid and amino acid sequences of the ScPIP gene family presented in this study.

Additional file 2: Table S2.

The bioinformatics information of ScPIP proteins.

Additional file 3: Table S3.

Primers used in this study.

Additional file 4: Table S4.

The gene amplification efficiency of RT-qPCR in this study.

Additional file 5: Fig. S1.

The standard curve plot of RT-qPCR primers in this study. Five points of a 10-fold dilution series of cDNA, performed in triplicate wells, amplified using the ABI 7500 real-time PCR system. The standard curve was generated by plotting threshold cycle (Ct) values against relative input cDNA dilution ratio. Taking logarithm value of cDNA dilution ratio as X-axis, ΔCt of gene as Y-axis. The cDNA of ROC22 leaf tissue was used for the standard curve analysis of the ScPIP2–1 and ScGAPDH genes. The cDNA of wild-type Arabidopsis Col-0 leaf tissue was used for the standard curve analysis of the AtKIN2, AtTRX5, AtP5CS1, AtP5CS2, AtDREB2, AtRD29A, AtNHX1, AtHKT1, AtSOS1 and AtACT genes.

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

Tang, H., Yu, Q., Li, Z. et al. A PIP-mediated osmotic stress signaling cascade plays a positive role in the salt tolerance of sugarcane. BMC Plant Biol 21, 589 (2021). https://doi.org/10.1186/s12870-021-03369-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12870-021-03369-9

Keywords