Next Article in Journal
UHPLC-Q-TOF/MS-Based Metabolomics Approach Reveals the Antifungal Potential of Pinocembroside against Citrus Green Mold Phytopathogen
Next Article in Special Issue
Role of Proton Motive Force in Photoinduction of Cytoplasmic Streaming in Vallisneria Mesophyll Cells
Previous Article in Journal
Copy Number Variation of Transposable Elements in Thinopyrum intermedium and Its Diploid Relative Species
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Expression and Ion Transport Activity of Rice OsHKT1;1 Variants

1
Institute of Plant Science and Resources, Okayama University, 2-20-1, Chuo, Kurashiki 710-0046, Japan
2
Division of Applied Biology, Faculty of the Textile Science and Technology, Shinshu University, 3-15-1, Tokida, Ueda 386-8567, Japan
*
Authors to whom correspondence should be addressed.
Plants 2020, 9(1), 16; https://doi.org/10.3390/plants9010016
Submission received: 21 November 2019 / Revised: 11 December 2019 / Accepted: 19 December 2019 / Published: 21 December 2019
(This article belongs to the Special Issue Plant Membrane Transporters)

Abstract

:
OsHKT1;1 in rice, belongs to the high-affinity K+ Transporter family, has been found to be involved in salt tolerance. OsHKT1;1 in japonica rice (Nipponbare) produces mRNA variants, but their functions remain elusive. In salt tolerant rice, Pokkali, eight OsHKT1;1 variants (V1-V8) were identified in addition to the full-length OsHKT1;1 (FL) cDNA. Absolute quantification by qPCR revealed that accumulation of OsHKT1;1-FL mRNA is minor in contrast to that of OsHKT1;1-V1, -V2, -V4, and -V7 mRNAs, all of which are predominant in shoots, while only V1 and V7 mRNAs are predominant in roots. Two electrode voltage clamp (TEVC) experiments using Xenopus laevis oocytes revealed that oocytes-expressing OsHKT1;1-FL from Pokkali exhibited inward-rectified currents in the presence of 96 mM Na+ as reported previously. Further TEVC analyses indicated that six of eight OsHKT1;1 variants elicited currents in a Na+ or a K+ bath solution. OsHKT1;1-V6 exhibited a similar inward rectification to the FL protein. Contrastingly, however, the rests mediated bidirectional currents in both Na+ and K+ bath solutions. These data suggest possibilities that novel mechanisms regulating the transport activity of OsHKT1;1 might exist, and that OsHKT1;1 variants might also carry out distinct physiological roles either independently or in combination with OsHKT1;1-FL.

1. Introduction

Salinity affects plant growth and development and causes low crop productivity [1]. Salinity can give rise to osmotic stress, nutritional imbalance, and ionic toxicity in the plant [2,3]. Among major cereal crops, rice (Oryza sativa) is the most sensitive to salinity stress [2,4].
The high-affinity K+ transporter (HKT) gene family has been found to be relevant to salt stress tolerance in plants [5,6]. The first HKT gene in plants, TaHKT2;1 from bread wheat Triticum aestivum, was reported to be a Na+-K+ co-transporter [7]. HKTs can be classified into at least two subfamilies [5,8]. Subfamily 1 HKTs (HKT1s) are found in monocotyledonous and dicotyledonous species, and they are Na+ selective transporters. So far, only monocotyledonous species contain the subfamily 2 HKTs (HKT2s) that are selective for both Na+ and K+ [9]. Interestingly, both HKT1 and HKT2 proteins were suggested to exhibit features of a Na+ channel when the Na+ concentration at extracellular spaces are high [7,10,11]. In fact, AtHKT1;1 in Arabidopsis, belongs to the subfamily 1, has been demonstrated to show a Na+ channel activity by comparative patch clamping analyses using wild-type and athkt1;1 loss-of-function mutants [12]. In a japonica cultivar of rice, seven functional OsHKT genes have been identified [13]. Four of them were grouped in the HKT1 gene subfamily and those products, OsHKT1;1, OsHKT1;3, OsHKT1;4 and OsHKT1;5, were shown to be Na+ selective by the two-electrode voltage clamp experiment using Xenopus laevis oocytes [14,15,16,17].
A QTL (Quantitative Trait Locus) analysis using a salt tolerance indica landrace, Nona Bokra, and a japonica cultivar, Koshihikari, has led to the identification of the shoot K+ content 1 (SKC1) locus, which regulates accumulation of higher K+ and lower Na+ in shoots under salt stress [14]. The OsHKT1;5 gene encoding a plasma membrane-localized Na+ transporter was suggested to be a causal gene for the QTL [14]. The SKC1-dependent phenotype in the accumulation of K+ and Na+ was similar to the mechanism found in the athkt1;1 mutant of Arabidopsis [18,19], which supports a hypothetical model that HKT1-mediated Na+ unloading at the xylem indirectly couples with K+ loading, which eventually accelerates lower Na+ but higher K+ accumulation in shoots [5]. More recently, OsHKT1;5 was demonstrated to have a critical role in reducing Na+ contents of young leaves of rice under salt stress via Na+ unloading at the xylem of roots and leaf sheaths and the phloem of the basal node [20].
Characterization of a Tos17 mutant rice line, which harbors an insertion of a transposon in the OsHKT1;1 gene has revealed that OsHKT1;1 plays an important role in reducing Na+ accumulation in the shoot to cope with salt stress [21]. The Tos17-insertion in OsHKT1;1 was found to cause a reduction in Na+ contents in the phloem sap but an increase in the xylem sap in response to salt stress, similar to athkt1;1 mutants [18,21]. In addition, the OsHKT1;1 gene has been strongly suggested to be a determinant of salt tolerance of rice by two independent genetic studies [22,23]. In particular, Campbell et al. [23] indicated that an OsHKT1;1 isoform identified in indica accessions exhibits higher inward Na+ currents than those of the japonica-predominant isoform. Such a difference was brought about due to a less negative voltage threshold of inward current activation of the indica-predominant isoform, which could account for higher root Na+ contents and superior salt stress tolerance of the indica accessions. In barley, allelic variation in the HvHKT1 gene is associated with higher salt tolerance of the Tibetan wild barley accessions in comparison with barley cultivars [24]. The HvHKT1 gene was later found to correspond to HKT1;1 by a phylogenetic analysis and demonstrated to encode a Na+ selective transporter [25]. Knocking-down of HvHKT1;1, mediated by barley stripe mosaic virus (BSMV)-induced gene silencing, increased sensitivity to salt stress and accumulation of Na+ in both roots and leaves in the barley lines compared with a vector-introduced control line, which demonstrated essentiality of this gene in the mechanism of salt tolerance of barley [25].
According to the gene database of the National Center for Biotechnology Information (NCBI), the OsHKT1;1 gene produces several mRNA variants in japonica rice group. Although evidence of an importance of the HKT1;1 gene for the salt tolerance mechanism has accumulated as mentioned above, functions and impacts on salt tolerance of these variants are not yet initiated. In this study, we have attempted to identify OsHKT1;1 variants in a salt tolerant landrace of indica rice, Pokkali, and characterized their expression patterns and transport properties by means of quantitative PCR (qPCR) analyses and two electrode voltage clamp experiments using X. laevis oocytes. New aspects of transport activity and expression profiles of OsHKT1;1 variants will be discussed.

2. Results

2.1. Isolation and Characterization of OsHKT1;1 cDNAs

Upon the cDNA isolation using primers for the full-length clone, several different fragments were amplified from the first strand population prepared from the whole seedling of a salt tolerant indica variety Pokkali (Figure 1A). Sequence analysis indicated the presence of eight splicing variants in addition to the full-length sequence of the OsHKT1;1 cDNA. The full length OsHKT1;1 clone (OsHKT1;1-FL) is constituted of 1831 nucleotides, which encodes 552 putative amino acid residues. The deduced amino acid sequence was 100% identical to OsHKT1;1 of a variety Zhenshan 2 [23]. Prediction of the structure of OsHKT1;1-FL by the online software TMHMM Server v. 2.0 resulted in nine transmembrane domains as shown in Populus PeHKT1;1 [26] (Figure 1B). Eight OsHKT1;1 variants, OsHKT1;1-V1, -V2, -V3, -V4, -V5, -V6, -V7, and -V8 cDNAs, were respectively composed of 1716, 1571, 1452, 1181, 1067, 958, 920, and 931 nucleotides, which encode 515, 318, 209, 217, 180, 223, 186, and 176 putative amino acid residues (Figure 1C, Figures S3 and S4).

2.2. Expression Analyses of OsHKT1;1 in Shoots and Roots

14 days old Pokkali plants were analyzed under a normal growth condition. qPCR analyses revealed that all OsHKT1;1 transcripts including the FL clone were detected in both roots and shoots, and OsHKT1;1-V1 was the most abundant variant among them (Figure 2). Results also showed that transcript levels of OsHKT1;1-V1, -V2, and -V7 were significantly higher than the rests with a tendency of higher accumulation of -V4 transcripts in shoots. In contrast, OsHKT1;1-V1 and -V7 transcripts were more abundant than the others in roots (Figure 2).
We next investigated the expression of OsHKT1;1 under salt stress. Time dependent effects of 100 mM NaCl treatments were analyzed by qPCR using samples from shoots and roots of Pokkali plants (Figure 3). Expression of the OsHKT1;1-FL transcript in shoots was transiently reduced at 24 h and again upregulated at 48 h of salt stress (Figure 3A). In contrast, its expression in roots was constantly lower (Figure 3A). Among the variants that highly accumulate in shoots, only OsHKT1;1-V1 showed an increase trend in accordance with the time after salt stress (Figure 3B). On the other hands, OsHKT1;1-V2 and -V4 showed a similar trend such that the transcript levels decreased up to 24 h and then increased at 48 h after the stress treatment (Figure 3C,E). OsHKT1;1-V7 transcripts did not show any characteristic difference in the pattern of its accumulation (Figure 3H). In roots, some fluctuation was found but no notable pattern was observed in most of the transcripts except for OsHKT1;1-V6, which showed a significant reduction in the accumulation in response to salt stress (Figure 3G).

2.3. Ion Transport Activity of OsHKT1;1-FL and Its Variants

X. laevis oocytes were injected with cRNAs of OsHKT1;1-FL and its variants, and two electrode voltage clamp (TEVC) experiments were performed. The results exhibited marked differences in ion transport properties of OsHKT1;1 transporters in Na+- or K+-containing solutions (Figure 4). As has been reported previously [15], OsHKT1;1-FL mediated inward rectifying currents in the presence of 96 mM Na+ (Figure 4A). We found that OsHKT1;1-V6 also showed a similar inward rectifying feature in the same solution although elicited currents were smaller than OsHKT1;1-FL (Figure 4A). OsHKT1;1-V1 and -V4 showed small currents similar to water-injected control oocytes (Figure 4A). In contrast, OsHKT1;1-V2, -V3, -V5, -V7, -V8 mediated Na+ currents in a bidirectional transport manner depending on the membrane voltage (Figure 4B).
Ion transport activity of OsHKT1;1 transporters was also investigated in the presence of 96 mM K+. Only small currents comparable to water-injected oocytes were observed in oocytes expressing OsHKT1;1-FL, -V1 and -V4 (Figure 4C). In OsHKT1;1-V6-expressing oocytes, weak currents, which were slightly higher than water-injected control, were observed with weak rectification. (Figure 4C). Interestingly, expression of OsHKT1;1-V2, -V3, -V5, -V7, and -V8 elicited more robust currents in the K+ bath solution (Figure 4D).

3. Discussion

Under salt stress several Na+ transporters play an essential role in Na+ tolerance in plants. Among them, HKT1s have been shown to have a crucial role in both mono and dicotyledonous plants in the tolerance to salinity stress [3,27].
Four distinct sequences of OsHKT1;1 transcripts in japonica rice have been reported (cation transporter HKT4-like: https://www.ncbi.nlm.nih.gov/gene/?term=LOC9266695). In the present study, several splice variants of OsHKT1;1 were detected in a salt-tolerant indica rice Pokkali (Figure 1). Among the previously registered OsHKT1;1 transcripts from japonica rice, the XM_015778164.2 (isoform X1) is the most identical to OsHKT1;1-FL isolated from Pokkali in the present study and corresponds to the japonica OsHKT1;1 clone, the product of which has been well characterized by TEVC experiments [15,23]. In fact, OsHKT1;1-FL from Pokkali was 100% identical to the OsHKT1;1 isoform found in a variety Zhenshan 2, which has been suggested to contribute to conferring more salt tolerance to indica rice accessions [23]. The XM_015778165.2 (isoform X2) and the XM_01577817.1 (isoformX4) were respectively the most identical to OsHKT1;1-V1 and -V2 clones in the present study (Figure 1C). The XM_015778166.2 (isoform X3), however, did not match any variants isolated in the present study. Therefore, cDNA clones of OsHKT1;1-V3, -V4, -V5, -V6, -V7, and -V8 were found only in Pokkali thus far.
Na+ exclusion mechanisms mediated by HKT1 transporters has been well studied in several plant species. Arabidopsis has only one HKT gene, AtHKT1;1, which was demonstrated to show passive Na+ transport properties under salt stress [11,12]. AtHKT1;1 is mainly expressed in xylem parenchyma cells and AtHKT1;1-mediated unloading of Na+ at the xylem was proposed to be a major physiological role of AtHKT1;1 in the mechanism of salt tolerance in Arabidopsis [5,18]. In addition, expression of AtHKT1;1 in the phloem of leaves was also reported [28] and AtHKT1;1-involved Na+ recirculation from shoots to roots via the phloem, which contributes to reducing Na+ accumulation in leaves and thus tolerance to salt stress, was also proposed [9,18,28]. Salt tolerance QTL analyses using durum wheat have revealed that Na+ exclusion from leaves is an essential process to cope with salt stress and governed by two loci Nax1 and Nax2 (Na+ exclusion 1 and 2) [29,30]. Subsequent studies suggested that causal genes for Nax1 and Nax2 loci are respectively TmHKT1;4-A2 and TmHKT1;5, both of which encode a Na+ selective transporter [1,9,31,32]. Evidence for important contribution of HKT1;4 and HKT1;5 genes in rice to salt stress tolerance also accumulated. OsHKT1;5 encoding a Na+ selective transporter, which was suggested to be a causal gene for the SKC1 (Shoot K+ Content 1) locus identified in a salt tolerance QTL analysis of rice, has been indicated to protect leaves including the newest expanding one presumably through Na+ unloading from the xylem of roots and sheaths and the phloem at the basal node under salt stress [14,15]. A Na+ selective transporter OsHKT1;4 in rice was proposed to function in Na+ exclusion from leaves and stems under salt stress at the reproductive growth stage [17]. In addition to these OsHKT1 genes, OsHKT1;1 has been reported to play a crucial role in Na+ exclusion from shoots in rice [21]. Predominant expression of OsHKT1;1 in the phloem evokes possible involvement of OsHKT1;1 in the Na+ recirculation mechanism from shoots to roots [21], however, detailed physiological functions of OsHKT1;1 is yet to be elucidated. Furthermore, more recently, a genetic study using diverse rice accessions provided evidence that indica accessions exhibit higher Na+ contents in roots than those in japonica accessions upon a moderate salt stress, which is regulated by the locus Root Na+ Content 4 (RNC4) [23]. The study further indicated that a strong candidate for the causal gene of the RNC4 appears to be OsHKT1;1 and that an OsHKT1;1 isoform found in indica accessions exhibits a higher Na+ transport activity with a less negative voltage threshold of inward-rectifying current activation in comparison with OsHKT1;1 in japonica accessions. Based on the observation, the study suggested that such a difference in the function of the OsHKT1;1 isoform contribute to higher salt tolerance of indica accessions.
As a matter of fact, OsHKT1;1-FL identified in this study was indeed 100% identical to the OsHKT1;1 isoform identified by Campbell et al. [19,23]. However, functions and roles of OsHKT1;1 variants have not been investigated so far. As shown in Figure 4A,B, OsHKT1;1-V2, -V3, -V5, -V6, -V7, and -V8, which were expressed in X. laevis oocytes, interestingly, elicited more robust currents in comparison with water-injected control oocytes in the presence of 96 mM Na+. Surprisingly, those variants also mediated robust currents in a bath solution containing 96 mM K+ (Figure 4C, D). The selectivity of HKT transporters for Na+ and K+ has been reported to be largely dependent on a key amino acid residue, serine (S) or glycine (G), at the key position in the putative pore-forming region (p-loop) ([9] and references therein). The key amino acid position is relevant to the GYG motif of a shaker K+ channel, which is an essential factor for the K+ selectivity ([9] and references therein). HKT transporters in general retain four p-loop regions and either S or G in the first p-loop can be consistent with its transport property, that is, a Na+ selective type (subfamily I) or a Na+-K+ co-transport type (subfamily II), respectively [5,9]. OsHKT1;1 retains the S residue at the key position of the first p-loop, which lies in the position of 104–121 amino acids of OsHKT1;1-FL. Except for OsHKT1;1-V2 that retains the rest of three p-loop regions, all other variants kept the first p-loop. Elucidating how those shorter variants can form a mature ion transporter and these subfamily 1-derived variants possibly mediates K+ transport will be an important subject to address in the future.
Wang et al. [21] reported approximately 3- to 5-fold increases in the OsHKT1;1 expression in 12 h after a salt stress in shoots, while no difference in its expression was observed in roots. In the present study, the level of OsHKT1;1-FL mRNA did not show a significant difference, but only OsHKT1;1-V1 mRNA significantly increased in shoots 12 h after the imposition of salt stress (Figure 3). Note however that, looking at 48 h after salt stress, levels of OsHKT1;1-V2, -V4, -V5, and -V8 mRNA were significantly increased (Figure 3). Our qPCR analyses also revealed that the OsHKT1;1-V1 mRNA was the most abundant in both roots and shoots among the variants identified (Figure 2). However, V1 showed no transport activity for Na+ and K+ in X. laevis oocytes (Figure 4). Increases in the OsHKT1;1-V1 mRNA after salt stress in shoots (Figure 3B) might have competitive effects on the function of other variants/FL. In addition, OsHKT1;1-V2 and -V7 might also be an interesting target for further research as both variants mediated bidirectional currents stimulated by Na+-and K+ in oocytes and their expression levels in Pokkali tissues were relatively high compared with OsHKT1;1-FL (Figure 2, Figure 3, and Figure 4). These results imply that not only OsHKT1;1-FL but also other variants could carry out some role in Pokkali plants under salt stress.
Pokkali OsHKT1;1-FL in the present study showed an inward-rectifying characteristic in Na+ transport as previously reported [15,23]. In contrast, OsHKT1;1-V1 showed no ion transport activity probably because of the loss of the M1 domain from the full-length sequence. However, the function of M1 seems not to be simple as OsHKT1:1-V2 that lacks both M1 and M2 domains showed ion transport activity (Figure 4), suggesting a role of M2 in cancelling the negative influence of the lack of M1 on the activity. Shorter variants suggest complex interactions between M1 and M2 domains and/or latter domains. It should be noted however that possibilities of different rates of translation and/or protein turnover among OsHKT1;1 variants were not taken into account in electrophysiological characterization by TEVC experiments in this study. Therefore, such aspects need to be considered in future TEVC experiments that will be performed to characterize OsHKT1;1 variants in more details. Molecular mechanisms of domain interactions and regulation in the ion transport activity and ion selectivity should also be revealed in future studies.

4. Materials and Methods

4.1. Plant Material and Growth Condition

A salt tolerant landrace of rice, Pokkali, (Oryza sativa L. ssp. indica) was used in this study. Seeds were sterilized two times by using 50% bleach for 10 min and then washed with sterilized water for 5 times. For germination, ten seeds were placed on a mesh in a sterilized pot with 20 mL of 1 mM CaSO4. Plants were grown in a controlled environment growth chamber. The temperature was maintained at 28 °C and 25 °C in 12 h day (250 μmol m−2 s−1 illumination) and 12 h night, respectively. Five-day-old seedlings were transferred to 3.5-liter pots and grown hydroponically with the solutions containing 4 mM KNO3, 1 mM NH4H2PO4.2H2O, 1 mM CaCl2.2H2O, 1 mM MgSO4.7H2O, and micronutrients (1 ppm Fe, 0,5 ppm B, 0.5 ppm Mn, 0.05 ppm Zn, 0.02 ppm Cu, and 0.01 ppm Mo). The pH of the nutrient solution was adjusted to 5.5 with NaOH. 9-day old plants were used for cDNA isolation. For the gene expression study, 14 day-old plants were subjected to 100 mM NaCl stress for 0 h (control), 6 h, 12 h, 24 h, and 48 h. Roots and shoots were then collected separately at different time points.

4.2. Extraction of DNA and RNA, and cDNA Synthesis

Pokkali genomic DNA was extracted from young leaves using a Cica Geneus DNA Prep Kit for Plants (Kanto, Japan) following the manufacturer’s instructions. Total RNA was extracted from root or shoot samples using a RNeasy Plant Mini Kit (Qiagen, Hilden, Germany) and then treated with RNase-free DNase I (Ambion, Austin, TX, USA). The quality and integrity of RNA samples were determined by a NanoDrop ND-1000 Spectrophotometer (Nanodrop, Wilmington, DE, USA). The first strand coding DNA (cDNA) was synthesized from total RNA using ReverTra Ace qPCR RT Master Mix (TOYOBO, Osaka City, Japan), according to the manufacturer’s instructions. OsHKT1;1 cDNAs was amplified by using OsHKT1;1 cloning primers (Supplemental Table S1) and confirmed by analyzing the sequence after cloning into the pCR4 topo vector (Invitrogen, Carlsbad, CA, USA).

4.3. Secondly Structure Analysis

Transmembrane domains were predicted by the online software TMHMM Server v. 2.0 (http://www.cbs.dtu.dk/services/TMHMM/) applied to Populus PeHKT1;1 [26].

4.4. Expression Analysis

Total RNA samples were isolated from shoot and root tissues of 14 day-old plants (control and 100 mM NaCl) and the first-strand cDNA was reverse transcribed with a high capacity cDNA reverse transcription kit (Applied Biosystem, Foster City, CA, USA). The OsHKT1;1 variants were amplified using specific primers (Supplement Table S1). Amplification of the actin mRNA was used as an internal control as described previously [33]. Absolute quantification was performed in the quantitative real-time PCR analysis using the 7300 real-time PCR machine (Applied Biosystem, Foster City, CA, USA) as described in [34] to analyze the expression level of each OsHKT1;1 variant. Specific cDNAs were used as a standard to quantify each variant, and the primer information was indicated in the Supplementary Figure S1.

4.5. Expression in Xenopus laevis (X. leavis) Oocytes

OsHKT1;1 cDNAs from Pokkali were subcloned into the pXβG vector between the 5´ and 3´ untranslated regions of the Xenopus β-globin gene. Capped cRNAs were synthesized in vitro from the linearized vector using the mMESSAGE mMACHINE T3 kit (Ambion, Austin, TX, USA). Oocytes were isolated and injected with 50 ng/50 nL of OsHKT1;1 cRNA solutions or with 50 nL of nuclease-free water (for control oocytes), which were then incubated at 18 °C in modified Barth’s solution (MBS; 88 mM NaCl, 1 mM KCl, 2.4 mM NaHCO3, 15 mM Tris-HCl (pH 7.6), 0.3 mM Ca(NO3)2.4H2O, 0.41 mM CaCl2.4H2O, 0.82 mM MgSO4.7H2O, 10 μg ml−1 sodium penicillin, and 10 μg ml−1 streptomycin sulfate) until electrophysiological recordings as described previously [35].

4.6. Electrophysiology

Whole oocyte currents were recorded using the two-electrode voltage-clamp (TEVC) technique 1 to 2 d after the cRNA injection. TEVC recordings and data analysis were performed using an Axoclamp 900A amplifier, an Axon Instruments Digidata 1440A, Clampex 10.3, and Clampfit 10.3 (Molecular Devices, San Jose, CA, USA). All bath solutions contained a background of 1.8 mM MgCl2, 1.8 mM CaCl2, and 10 mM HEPES pH 7.5 with Tris. Monovalent cation was used as chloride salts (chloride concentration constant in each set of solutions). D-mannitol was added when necessary to adjust the osmolality (same osmolality in each set of solutions in the range 200–220 mOsm). Voltage steps (2 sec) were applied from +30 to −135 mV in 15 mV decrements. All experiments were performed at room temperature (20–22 °C).

4.7. Statistics

Statistical analyses were performed using IBM SPSS Statistics Desktop for Japan. Significant differences were identified by one-way analysis of variance followed by Tukey HSD (p < 0.05).

Supplementary Materials

The following are available online at https://www.mdpi.com/2223-7747/9/1/16/s1, Table S1: Gene-specific primer pairs used for the cloning and in the real-time PCR experiments, Figure S1: Specific primer position of OsHKT1;1-FL and its eight variant for qPCR assay, Figure S2: Nucleotide sequence alignment of Pokkali OsHKT1;1 genomic DNA and full-length cDNA, Figure S3: Nucleotide sequence alignment of OsHKT1;1-FL and its eight variant, Figure S4; Protein sequence alignment of OsHKT1;1-FL and its eight variant, Figure S5: Expression analyzed with RT-PCR.

Author Contributions

Conceptualization, M.K. and T.H.; investigation, T.H. and S.I.; data analysis, M.K. and S.I.; writing―original draft preparing, S.I.; writing―review and editing, M.K. and T.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Ohara Foundation for Agricultural Science, (to M.K.), and by the Ministry of Education, Culture, Sports, Science and Technology (MEXT) as part of the Joint Research Program implemented at the Institute of Plant Science and Resources, Okayama University in Japan (2918, 3019 to T.H.).

Acknowledgments

We thank Yoshiyuki Tsuchiya for experimental assistance.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.

References

  1. Munns, R.; James, R.A.; Xu, B.; Athman, A.; Conn, S.J.; Jordans, C.; Byrt, C.S.; Hare, R.A.; Tyerman, S.D.; Tester, M.; et al. Wheat grain yield on saline soils is improved by an ancestral Na+ transporter gene. Nat. Biotechnol. 2012, 30, 360–364. [Google Scholar] [CrossRef] [PubMed]
  2. Munns, R.; Tester, M. Mechanisms of salinity tolerance. Annu. Rev. Plant Biol. 2008, 59, 651–681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Ismail, A.M.; Horie, T. Genomics, physiology, and molecular breeding approaches for improving salt tolerance. Annu. Rev. Plant Biol. 2017, 68, 405–434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. United States Department of Agriculture. Bibliography on Salt Tolerance. Fibres, Grains and Special Crops; Brown, G.E., Jr., Ed.; Salinity Laboratory United States Department of Agriculture Research Service: Riverside, CA, USA. Available online: http://www.Ars.Usda.Gov/services/docs.Htm?Docid=8908 (accessed on 20 August 2016).
  5. Horie, T.; Hauser, F.; Schroeder, J.I. HKT transporter-mediated salinity resistance mechanisms in Arabidopsis and monocot crop plants. Trends Plant Sci. 2009, 14, 660–668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Rubio, F.; Nieves-Cordones, M.; Horie, T.; Shabala, S. Doing ‘business as usual’ comes with a cost: Evaluating energy cost of maintaining plant intracellular K+ homeostasis under saline conditions. New Phytol. 2019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Rubio, F.; Gassmann, W.; Schroeder, J.I. Sodium-driven potassium uptake by the plant potassium transporter HKT1 and mutations conferring salt tolerance. Science 1995, 270, 1660–1663. [Google Scholar] [CrossRef]
  8. Platten, J.D.; Cotsaftis, O.; Berthomieu, P.; Bohnert, H.; Davenport, R.J.; Fairbairn, D.J.; Horie, T.; Leigh, R.A.; Lin, H.X.; Luan, S.; et al. Nomenclature for HKT transporters, key determinants of plant salinity tolerance. Trends Plant Sci. 2006, 11, 372–374. [Google Scholar] [CrossRef]
  9. Hauser, F.; Horie, T. A conserved primary salt tolerance mechanism mediated by HKT transporters: A mechanism for sodium exclusion and maintenance of high K+/Na+ ratio in leaves during salinity stress. Plant Cell Environ. 2010, 33, 552–565. [Google Scholar] [CrossRef]
  10. Gassmann, W.; Rubio, F.; Schroeder, J.I. Alkali cation selectivity of the wheat root high-affinity potassium transporter HKT1. Plant J. 1996, 10, 869–882. [Google Scholar] [CrossRef]
  11. Uozumi, N.; Kim, E.J.; Rubio, F.; Yamaguchi, T.; Muto, S.; Tsuboi, A.; Bakker, E.P.; Nakamura, T.; Schroeder, J.I. The Arabidopsis HKT1 gene homolog mediates inward Na+ currents in Xenopus laevis oocytes and Na+ uptake in Saccharomyces cerevisiae. Plant Physiol. 2000, 122, 1249–1259. [Google Scholar] [CrossRef] [Green Version]
  12. Xue, S.; Yao, X.; Luo, W.; Jha, D.; Tester, M.; Horie, T.; Schroeder, J.I. AtHKT1;1 Mediates Nernstian Sodium Channel Transport Properties in Arabidopsis Root Stellar Cells. PLoS ONE 2011, 6, e24725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Garciadeblás, B.; Senn, M.E.; Bañuelos, M.A.; Rodríguez-Navarro, A. Sodium transport and HKT transporters: The rice model. Plant J. 2003, 34, 788–801. [Google Scholar] [CrossRef] [PubMed]
  14. Ren, Z.H.; Gao, J.P.; Li, L.G.; Cai, X.L.; Huang, W.; Chao, D.Y.; Zhu, M.Z.; Wang, Z.Y.; Luan, S.; Lin, H.X. A rice quantitative trait locus for salt tolerance encodes a sodium transporter. Nat. Genet. 2005, 37, 1141–1146. [Google Scholar] [CrossRef] [PubMed]
  15. Jabnoune, M.; Espeout, S.; Mieulet, D.; Fizames, C.; Verdeil, J.L.; Conéjéro, G.; Rodríguez-Navarro, A.; Sentenac, H.; Guiderdoni, E.; Abdelly, C.; et al. Diversity in expression patterns and functional properties in the rice HKT transporter family. Plant Physiol. 2009, 150, 1955–1971. [Google Scholar] [CrossRef] [Green Version]
  16. Rosas-Santiago, P.; Lagunas-Gómez, D.; Barkla, B.J.; Vera-Estrella, R.; Lalonde, S.; Jones, A.; Frommer, W.B.; Zimmermannova, O.; Sychrová, H.; Pantoja, O. Identification of rice cornichon as a possible cargo receptor for the Golgi-localized sodium transporter OsHKT1;3. J. Exp. Bot. 2015, 66, 2733–2748. [Google Scholar] [CrossRef]
  17. Suzuki, K.; Costa, A.; Nakayama, H.; Katsuhara, M.; Shinmyo, A.; Horie, T. OsHKT2;2/1‑mediated Na+ influx over K+ uptake in roots potentially increases toxic Na+ accumulation in a salt-tolerant landrace of rice Nona Bokra upon salinity stress. J. Plant Res. 2016, 129, 67–77. [Google Scholar] [CrossRef]
  18. Sunarpi, H.T.; Horie, T.; Motoda, J.; Kubo, M.; Yang, H.; Yoda, K.; Horie, R.; Chan, W.Y.; Leung, H.Y.; Hattori, K.; et al. Enhanced salt tolerance mediated by AtHKT1 transporter-induced Na unloading from xylem vessels to xylem parenchyma cells. Plant J. 2005, 44, 928–938. [Google Scholar] [CrossRef]
  19. Møller, I.S.; Gilliham, M.; Jha, D.; Mayo, G.M.; Roy, S.J.; Coates, J.C.; Haseloff, J.; Tester, M. Shoot Na+ exclusion and increased salinity tolerance engineered by cell type-specific alteration of Na+ transport in Arabidopsis. Plant Cell 2009, 21, 2163–2178. [Google Scholar] [CrossRef] [Green Version]
  20. Kobayashi, N.I.; Yamaji, N.; Yamamoto, H.; Okubo, K.; Ueno, H.; Costa, A.; Tanoi, K.; Matsumura, H.; Fujii-Kashino, M.; Horiuchi, T.; et al. OsHKT1;5 mediates Na+ exclusion in the vasculature to protect leaf blades and reproductive tissues from salt toxicity in rice. Plant J. 2017, 91, 657–670. [Google Scholar] [CrossRef] [Green Version]
  21. Wang, R.; Jing, W.; Xiao, L.; Jin, Y.; Shen, L.; Zhang, W. The Rice High-Affinity Potassium Transporter1;1 Is Involved in Salt Tolerance and Regulated by an MYB-Type Transcription Factor. Plant Physiol. 2015, 168, 1076–1090. [Google Scholar] [CrossRef] [Green Version]
  22. Takagi, H.; Tamiru, M.; Abe, A.; Yoshida, K.; Uemura, A.; Yaegashi, H.; Obara, T.; Oikawa, K.; Utsushi, H.; Kanzaki, E.; et al. MutMap accelerates breeding of a salt-tolerant rice cultivar. Nat. Biotechnol. 2015, 33, 445–449. [Google Scholar] [CrossRef] [PubMed]
  23. Campbell, M.T.; Bandillo, N.; Al Shiblawi, F.R.A.; Sharma, S.; Liu, K.; Du, Q.; Schmitz, A.J.; Zhang, C.; Very, A.A.; Lorenz, A.J.; et al. Allelic variants of OsHKT1;1 underlie the divergence between indica and japonica subspecies of rice (Oryza sativa) for root sodium content. PLoS Genet. 2017, 13, e1006823. [Google Scholar] [CrossRef] [PubMed]
  24. Qiu, L.; Wu, D.; Ali, S.; Cai, S.; Dai, F.; Jin, X.; Wu, F.; Zhang, G. Evaluation of salinity tolerance and analysis of allelic function of HvHKT1 and HvHKT2 in Tibetan wild barley. Theor. Appl. Genet. 2011, 122, 695–703. [Google Scholar] [CrossRef] [PubMed]
  25. Han, Y.; Yin, S.; Huang, L.; Wu, X.; Zeng, J.; Liu, X.; Qiu, L.; Munns, R.; Chen, Z.-H.; Zhang, G. A Sodium Transporter HvHKT1;1 Confers Salt Tolerance in Barley via Regulating Tissue and Cell Ion Homeostasis. Plant Cell Physiol. 2018, 59, 1976–1989. [Google Scholar] [CrossRef]
  26. Xu, M.; Chen, C.; Cai, H.; Wu, L. Overexpression of PeHKT1;1 Improves Salt Tolerance in Populus. Genes 2018, 9, 475. [Google Scholar] [CrossRef] [Green Version]
  27. Deinlein, U.; Stephan, A.B.; Horie, T.; Luo, W.; Xu, G.; Schroeder, J.I. Plant salt-tolerance mechanisms. Trends Plant Sci. 2014, 19, 371–379. [Google Scholar] [CrossRef] [Green Version]
  28. Berthomieu, P.; Conéjéro, G.; Nublat, A.; Brackenbury, W.J.; Lambert, C.; Savio, C.; Uozumi, N.; Oiki, S.; Yamada, K.; Cellier, F.; et al. Functional analysis of AtHKT1 in Arabidopsis shows that Na+ recirculation by the phloem is crucial for salt tolerance. EMBO J. 2003, 22, 2004–2014. [Google Scholar] [CrossRef] [Green Version]
  29. Munns, R.; Rebetzke, G.J.; Husain, S.; James, R.A.; Hare, R.A. Genetic control of sodium exclusion in durum wheat. Aust. J. Agric. Res. 2003, 54, 627–635. [Google Scholar] [CrossRef]
  30. James, R.A.; Davenport, R.J.; Munns, R. Physiological characterization of two genes for Na+ exclusion in durum wheat, Nax1 and Nax2. Plant Physiol. 2006, 142, 1537–1547. [Google Scholar] [CrossRef] [Green Version]
  31. Byrt, C.S.; Platten, J.D.; Spielmeyer, W.; James, R.A.; Lagudah, E.S.; Dennis, E.S.; Tester, M.; Munns, R. HKT1;5-like cation transporters linked to Na+ exclusion loci in wheat, Nax2 and Kna1. Plant Physiol. 2007, 143, 1918–1928. [Google Scholar] [CrossRef] [Green Version]
  32. Tounsi, S.; Amar, B.S.; Masmoudi, K.; Sentenac, H.; Brini, F.; Véry, A.A. Characterization of Two HKT1;4 Transporters from Triticum monococcum to Elucidate the Determinants of the Wheat Salt Tolerance Nax1 QTL. Plant Cell Physiol. 2016, 57, 2047–2057. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Katsuhara, M.; Rhee, J.; Sugimoto, G.; Chung, G.C. Early response in water relations influenced by NaCl reflects tolerance or sensitivity of barley plants to salinity stress via aquaporins. Soil Sci. Plant Nutr. 2010, 57, 50–60. [Google Scholar] [CrossRef]
  34. 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, 801–813. [Google Scholar] [CrossRef] [PubMed]
  35. Katsuhara, M.; Akiyama, Y.; Koshio, K.; Shibasaka, M.; Kasamo, K. Functional analysis of water channels in barley roots. Plant Cell Physiol. 2002, 43, 885–893. [Google Scholar] [CrossRef] [PubMed]
Figure 1. OsHKT1;1 transcripts identified in Pokkali plants. (A) A gel image of RT-PCR produces amplified with primers for OsHKT1;1; (B) the transmembrane domain of the full-length OsHKT1;1 protein (OsHKT1;1-FL) predicted by TMHMM Server v.2.0. Nine transmembrane domains (M1-M9) were predicted; (C) schematic diagrams of OsHKT1;1-FL and its variants. Boxes indicate amino acid regions that are the same as FL (black) or different from FL because of the frame shift (grey). The line “─” indicates non-translated regions, and dot line “┉” indicates missing nucleotide regions (gap) compared with the FL sequence.
Figure 1. OsHKT1;1 transcripts identified in Pokkali plants. (A) A gel image of RT-PCR produces amplified with primers for OsHKT1;1; (B) the transmembrane domain of the full-length OsHKT1;1 protein (OsHKT1;1-FL) predicted by TMHMM Server v.2.0. Nine transmembrane domains (M1-M9) were predicted; (C) schematic diagrams of OsHKT1;1-FL and its variants. Boxes indicate amino acid regions that are the same as FL (black) or different from FL because of the frame shift (grey). The line “─” indicates non-translated regions, and dot line “┉” indicates missing nucleotide regions (gap) compared with the FL sequence.
Plants 09 00016 g001
Figure 2. qPCR analyses on OsHKT1;1 transcripts in Pokkali plants grown under normal growth condition. Expression levels of OsHKT1;1-FL and its variants in shoots (A) and roots (B) of 14 day-old plants were investigated by absolute quantification. Insets; Expanded views of variants showing low expression. Data are means ± SE, n = 3 Three independent experiments were performed, and similar results were obtained. Different letters indicate significant differences (p < 0.05).
Figure 2. qPCR analyses on OsHKT1;1 transcripts in Pokkali plants grown under normal growth condition. Expression levels of OsHKT1;1-FL and its variants in shoots (A) and roots (B) of 14 day-old plants were investigated by absolute quantification. Insets; Expanded views of variants showing low expression. Data are means ± SE, n = 3 Three independent experiments were performed, and similar results were obtained. Different letters indicate significant differences (p < 0.05).
Plants 09 00016 g002
Figure 3. qPCR analyses on OsHKT1;1 transcripts in Pokkali seedlings grown under salt stressed conditions. Expression levels of OsHKT1;1-FL and each variant in shoots and roots were investigated by absolute quantification. 14 day-old Pokkali plants were treated with 100 mM NaCl solution for 0, 6, 12, 24, or 48 h prior to the total RNA extraction. (A) OsHKT1;1-FL, (B) OsHKT1;1-V1, (C) OsHKT1;1-V2, (D) OsHKT1;1-V3, (E) OsHKT1;1-V4, (F) OsHKT1;1-V5, (G) OsHKT1;1-V6, (H) OsHKT1;1-V7, (I) OsHKT1;1-V8. Absolute amounts of transcript (copies/μg RNA) were displayed. Data are means ± SE, n = 3. Three independent experiments were performed, and similar results were obtained. Different letters indicate significant differences (p < 0.05).
Figure 3. qPCR analyses on OsHKT1;1 transcripts in Pokkali seedlings grown under salt stressed conditions. Expression levels of OsHKT1;1-FL and each variant in shoots and roots were investigated by absolute quantification. 14 day-old Pokkali plants were treated with 100 mM NaCl solution for 0, 6, 12, 24, or 48 h prior to the total RNA extraction. (A) OsHKT1;1-FL, (B) OsHKT1;1-V1, (C) OsHKT1;1-V2, (D) OsHKT1;1-V3, (E) OsHKT1;1-V4, (F) OsHKT1;1-V5, (G) OsHKT1;1-V6, (H) OsHKT1;1-V7, (I) OsHKT1;1-V8. Absolute amounts of transcript (copies/μg RNA) were displayed. Data are means ± SE, n = 3. Three independent experiments were performed, and similar results were obtained. Different letters indicate significant differences (p < 0.05).
Plants 09 00016 g003
Figure 4. Ion transport activity of OsHKT1;1-FL and its variants. Two electrode voltage clamp experiments using Xenopus laevis oocytes were conducted. (A,C) Current-voltage relationships from oocytes expressing OsHKT1;1-FL, -V1, -V4, -V6, and water-injection control. (B,D) Current-voltage relationships from oocytes expressing OsHKT1;1-V2, -V3, -V5, -V7, -V8, and water-injection control. External solution contains 96 mM NaCl (A,B) or 96 mM KCl (B,D). All external solutions contain, as background elements, 1.8 mM CaCl2, 1.8 mM MgCl2, 1.8 mM mannitol, and 10 mM HEPES (pH 7.5 with Tris). Water was injected as a negative control. Data represent means ± SE, n = 18–24, from three independent experiments performed on different oocytes batches.
Figure 4. Ion transport activity of OsHKT1;1-FL and its variants. Two electrode voltage clamp experiments using Xenopus laevis oocytes were conducted. (A,C) Current-voltage relationships from oocytes expressing OsHKT1;1-FL, -V1, -V4, -V6, and water-injection control. (B,D) Current-voltage relationships from oocytes expressing OsHKT1;1-V2, -V3, -V5, -V7, -V8, and water-injection control. External solution contains 96 mM NaCl (A,B) or 96 mM KCl (B,D). All external solutions contain, as background elements, 1.8 mM CaCl2, 1.8 mM MgCl2, 1.8 mM mannitol, and 10 mM HEPES (pH 7.5 with Tris). Water was injected as a negative control. Data represent means ± SE, n = 18–24, from three independent experiments performed on different oocytes batches.
Plants 09 00016 g004

Share and Cite

MDPI and ACS Style

Imran, S.; Horie, T.; Katsuhara, M. Expression and Ion Transport Activity of Rice OsHKT1;1 Variants. Plants 2020, 9, 16. https://doi.org/10.3390/plants9010016

AMA Style

Imran S, Horie T, Katsuhara M. Expression and Ion Transport Activity of Rice OsHKT1;1 Variants. Plants. 2020; 9(1):16. https://doi.org/10.3390/plants9010016

Chicago/Turabian Style

Imran, Shahin, Tomoaki Horie, and Maki Katsuhara. 2020. "Expression and Ion Transport Activity of Rice OsHKT1;1 Variants" Plants 9, no. 1: 16. https://doi.org/10.3390/plants9010016

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop