Skip to main content
The Plant Cell logoLink to The Plant Cell
. 2008 Oct;20(10):2876–2893. doi: 10.1105/tpc.108.061374

Transgenic Arabidopsis Plants Expressing the Type 1 Inositol 5-Phosphatase Exhibit Increased Drought Tolerance and Altered Abscisic Acid Signaling[W]

Imara Y Perera a,1, Chiu-Yueh Hung a,2, Candace D Moore a,3, Jill Stevenson-Paulik b,4, Wendy F Boss a
PMCID: PMC2590728  PMID: 18849493

Abstract

The phosphoinositide pathway and inositol-1,4,5-trisphosphate (InsP3) are implicated in plant responses to stress. To determine the downstream consequences of altered InsP3-mediated signaling, we generated transgenic Arabidopsis thaliana plants expressing the mammalian type I inositol polyphosphate 5-phosphatase (InsP 5-ptase), which specifically hydrolyzes soluble inositol phosphates and terminates the signal. Rapid transient Ca2+ responses to a cold or salt stimulus were reduced by ∼30% in these transgenic plants. Drought stress studies revealed, surprisingly, that the InsP 5-ptase plants lost less water and exhibited increased drought tolerance. The onset of the drought stress was delayed in the transgenic plants, and abscisic acid (ABA) levels increased less than in the wild-type plants. Stomatal bioassays showed that transgenic guard cells were less responsive to the inhibition of opening by ABA but showed an increased sensitivity to ABA-induced closure. Transcript profiling revealed that the drought-inducible ABA-independent transcription factor DREB2A and a subset of DREB2A-regulated genes were basally upregulated in the InsP 5-ptase plants, suggesting that InsP3 is a negative regulator of these DREB2A-regulated genes. These results indicate that the drought tolerance of the InsP 5-ptase plants is mediated in part via a DREB2A-dependent pathway and that constitutive dampening of the InsP3 signal reveals unanticipated interconnections between signaling pathways.

INTRODUCTION

The membrane-associated phospholipids and the soluble inositol phosphates, collectively known as the phosphoinositides, are ubiquitous in eukaryotic cells and are involved in regulating a multitude of cellular functions (Martin, 1998; Cockcroft and De Matteis, 2001; Di Paolo and De Camilli, 2006). In the canonical phosphoinositide pathway, the membrane-associated phospholipid, phosphatidylinositol (PtdIns), is sequentially phosphorylated by specific lipid kinases to form phosphatidylinositol 4-phosphate (PtdInsP) and phosphatidylinositol 4,5 bisphosphate (PtdInsP2). In response to a stimulus such as hyperosmotic stress, PtdInsP2 is hydrolyzed by phospholipase C (PLC) to produce the soluble second messengers inositol 1,4,5-trisphosphate (InsP3) and diacylglycerol (Berridge, 1993). The phospholipids are ideally suited to mediate cellular signaling due to their high turnover rate and their spatial distribution, which is regulated in part by the concerted action of specific phosphoinositide kinases, phospholipases, and phosphatases (Mueller-Roeber and Pical, 2002; Boss et al., 2006). The phosphoinositide pathway integrates plasma membrane sensing with intracellular signaling and thereby provides a means of both sensing and propagating a signal.

Although the phosphoinositide pathway has been well studied in plants, there are fundamental differences between the plant and animal models (Stevenson et al., 2000; Meijer and Munnik, 2003; Wang, 2004). One critical difference is the relatively low steady state level of PtdInsP2 in plants compared with animals. One of the reasons for this difference is that the plant PtdInsP kinases are not as active as their animal counterparts (Perera et al., 2005; Im et al., 2007).

Although rapid, transient increases in InsP3 have been demonstrated in many different plant tissues in response to environmental stimuli (Stevenson et al., 2000; Meijer and Munnik, 2003; Krinke et al., 2007), the downstream consequences of the InsP3 changes are not well understood. InsP3 has been shown to trigger Ca2+ release from intracellular stores such as the vacuole (Sanders et al., 1999, 2002). InsP3 is implicated in the propagation of Ca2+ oscillations in guard cells and root hair cells (Staxen et al., 1999; Engstrom et al., 2002), in cell-to-cell communication (Tucker and Boss, 1996), and in the regulation of diurnal Ca2+ fluctuations (Tang et al., 2007). However, genes encoding a canonical InsP3 receptor are absent in the Arabidopsis thaliana genome (Hetherington and Brownlee, 2004; Krinke et al., 2007). In addition, only a few examples of a stimulus-generated increase in InsP3 resulting in a change in cytosolic Ca2+ and a downstream response have been reported in higher plants (Franklin-Tong et al., 1996; DeWald et al., 2001).

In order to manipulate the plant phosphoinositide pathway and understand the downstream effectors of InsP3-mediated signaling, we generated transgenic plants constitutively expressing the mammalian type I inositol polyphosphate 5-phosphatase (InsP 5-ptase), an enzyme that specifically hydrolyzes the soluble inositol phosphates InsP3 and inositol tetrakisphosphate (InsP4) and not the inositol phospholipids (Laxminarayan et al., 1993, 1994; Majerus et al., 1999). We chose this heterologous enzyme for several reasons. The mammalian type 1 inositol 5-phosphatase is well characterized and is more active than its plant counterparts. This enzyme is associated with the plasma membrane (De Smedt et al., 1997) and therefore should preferentially dampen InsP3 signals generated at the plasma membrane. In previous work, we have shown that this enzyme is stably expressed in plants and produces an active, functional protein that is plasma membrane–localized (Perera et al., 2002, 2006). A further advantage of the heterologous genes is that they are not subject to gene-silencing effects. In Arabidopsis, the inositol 5-phosphatases are encoded by a multigene family of 15 genes (Berdy et al., 2001; Ercetin and Gillaspy, 2004), which increases gene redundancy.

Our hypothesis was that stimulus-induced increases in InsP3 and InsP3-mediated signaling cascades would be impaired in the transgenic plants and thereby the timing and/or magnitude of the response would be delayed or decreased. In previous work, we have shown that InsP 5-ptase transgenic Arabidopsis plants have normal growth and morphology under optimal growth conditions; however, basal InsP3 levels are greatly reduced (to ∼5% of wild-type levels) and the plants exhibit delayed and reduced gravitropic responses in roots, hypocotyls, and inflorescence stems (Perera et al., 2006). These data support the hypothesis that InsP3 is a universal signal in plant gravitropism and that decreasing the signal attenuates the response.

In this work, we have used the InsP 5-ptase plants to dissect the function of InsP3 in drought stress. Abscisic acid (ABA) is a prime mediator of plant responses to abiotic stresses such as drought, salt, and cold (Zhu, 2002). In the drought stress response, ABA acts at multiple cellular levels, including the regulation of stomatal aperture and the activation of gene expression. Many second messengers, including Ca2+, InsP3, cADP ribose, and others, are implicated in ABA-mediated signaling. In addition, there is crosstalk between both ABA-dependent and ABA-independent pathways involving other drought-responsive factors such as DREB2A and ERD1 (Shinozaki and Yamaguchi-Shinozaki, 2000, 2007; Yamaguchi-Shinozaki and Shinozaki, 2006).

Both InsP3 and PLC (the enzyme responsible for the generation of InsP3 from PtdInsP2) have been shown to be important for ABA-mediated stomatal regulation. In early work in Commelina communis epidermal peels, the release of caged InsP3 caused an elevation of cytosolic Ca2+ leading to stomatal closure (Gilroy et al., 1990). Also, treatment with the PLC inhibitor U73122 inhibited ABA-induced Ca2+ oscillations and stomatal closure (Staxen et al., 1999). More recently, Hunt et al., 2003 reported that transgenic tobacco (Nicotiana tabacum) plants with reduced PLC in their guard cells showed a wilty phenotype under water-limiting conditions and did not fully respond to ABA. Further characterization of the ABA-mediated stomatal regulation in these plants (Mills et al., 2004) revealed that the ability of ABA to inhibit stomatal opening was compromised (in the reduced PLC plants), although there was no difference in ABA promotion of stomatal closure. Stomatal conductance was also higher in plants with reduced PLC after a period of soil drying compared with controls.

Ectopic expression of endogenous plant InsP 5-ptase genes has also revealed differences in ABA responses compared with the wild type. Stomata of transgenic plants overexpressing Arabidopsis InsP 5-ptase, At5PTase1, were shown to be less open than the wild type and less responsive to ABA (Burnette et al., 2003), with delayed induction of ABA-responsive gene expression. Induction of ABA-responsive genes was also attenuated in transgenic plants expressing another Arabidopsis inositol phosphatase, At5PTase2 (Sanchez and Chua, 2001).

Based on our predictions and the work by others (described above), we expected that the InsP 5-ptase transgenic plants would be less drought-tolerant and unable to withstand prolonged drought stress. Surprisingly, the transgenic plants lose less water and are more drought-tolerant than the wild-type plants. We show here that the guard cells of transgenic plants show differential sensitivity to exogenous ABA compared with wild-type plants in bioassays to monitor stomatal opening and closure. Furthermore, with drought stress, ABA levels in the transgenic plants do not increase as much as in the wild type. A key to understanding the drought tolerance phenotype is the greater than twofold increase in basal transcript levels of the drought-inducible ABA-independent transcription factor DREB2A and a subset of DREB2A-regulated genes in the unstressed transgenic plants. These results suggest that InsP3 acts as a negative regulator in the DREB2A drought signaling pathway and that in the absence of InsP3, this and other compensatory pathways have been activated in the transgenic plants, leading to increased drought tolerance.

RESULTS

The InsP 5-ptase Plants Show Increased Tolerance to Drought and Lose Less Water Compared with Wild-Type Plants

Several independent stably transformed Arabidopsis lines expressing InsP 5-ptase have been generated and characterized as described previously (Perera et al., 2006). The transgenic plants have no morphological changes compared with the wild type and grow comparably under normal growth conditions. Furthermore, the InsP 5-ptase protein is produced and active, as evidenced by immunoblotting and the fact that the basal InsP3 levels are reduced by >90% compared with wild-type plants (Perera et al., 2006).

In this work, we have studied the response of the transgenic plants to drought stress. When water is withheld, the transgenic plants are able to withstand the drought conditions longer than the wild-type plants. After 12 d without water, the wild-type plants and plants from the vector control line are wilted and turning brown, while plants from three independent InsP 5-ptase lines remain green and turgid (Figure 1A). We next monitored water loss from detached leaves using fully expanded leaves of comparable size, weight, and development. Leaves were excised from the rosettes of well-hydrated wild-type and transgenic plants and kept at ambient light and temperature at 30% humidity (adaxial side up) for 1 to 2 h, and the fresh weight of the leaves was monitored periodically during this time. The transgenic plants show a 30% decrease in the rate of water loss from detached leaves compared with the wild-type plants (Figure 1B). We also monitored pot water loss over a period of 6 to 7 d to measure water loss during day and night. Over this time period, the transgenic plants lost less water during both the day and night (Figure 1C). The extent of root growth and root mass between wild-type and transgenic plants as observed at the end the experiment (see Supplemental Figure 1 online) was comparable, ruling out the possibility that reduced water loss from the transgenic plants was caused by greater water retention due to differential root growth.

Figure 1.

Figure 1.

The InsP 5-ptase Transgenic Plants Lose Less Water and Are More Drought-Tolerant Compared with Wild-Type Plants.

(A) Eight-week-old plants were incubated at room temperature (30% humidity) and not watered for 12 d. C-2, vector control, hereafter denoted C2; 2-12, 2-6, and 2-8, three independent transgenic lines expressing InsP 5-ptase, hereafter denoted T12, T6, and T8, respectively.

(B) Leaf water loss in excised leaves was monitored at 5-min intervals over a period 1 to 2 h. Data are plotted as the decrease in fresh weight over time adjusted for potential differences in leaf surface area. Values are averages ± se of four independent experiments.

(C) Water loss of intact plants in pots during the day and night was measured over a period of 6 d. The data plotted are averages ± se from nine plants per line from either the wild type or two transgenic lines, T6 and T8, that were combined and denoted Tr.

(D) Stomatal conductance in well-watered plants measured over a period of 8 h. Data presented are averages ± se of three measurements per plant and six plants per line for each time point. The experiment was repeated twice with similar results.

(E) Stomatal conductance in drought-stressed plants (without water for 8 d) measured at two time points after being exposed to light. Data presented are averages ± se of three measurements per plant and 10 plants per line for each time point.

We also measured stomatal conductance in well-watered plants over a period of 6 to 8 h of exposure to light. During the time course, maximal stomatal conductance was detected in all plant lines after ∼4 h of light, which is consistent with the circadian rhythm of the plants. Stomatal conductance was comparable among the two independent transgenic plant lines and the wild-type and vector control plants (Figure 1D). These results demonstrate that the transgenic plants are drought tolerant and that under well-watered conditions stomatal conductance was not affected, consistent with their normal growth phenotype. However, after 8 d without water, stomatal conductance was reduced in all plant lines and, importantly, was lower in the transgenic plants compared with the wild type (Figure 1E). This suggests that the transgenic plants are better able to adjust to the water stress compared with the wild-type plants.

The Transgenic Guard Cells Show Altered Responses to ABA

Because stomata regulate the major flow of water out of leaves during both day and night (Caird et al., 2007), we examined the response of guard cells from wild-type and transgenic plants to different stimuli. For these experiments, epidermal peels were prepared from the abaxial surface of intact healthy leaves of similar size and developmental stage. The presence of the InsP 5-ptase protein (∼45 kD) in similar epidermal peel preparations was clearly visible by immunoblotting (see Supplemental Figure 2 online). InsP3 levels were reduced by 95% in epidermal peel preparations (Table 1), as in all other tissues of the transgenic plants tested (Perera et al., 2006). We also monitored stomatal density in epidermal peels. There was no difference in the number of stomata/leaf area between wild-type and transgenic leaves (Table 1).

Table 1.

InsP3 Levels and Stomatal Density in Epidermal Peels

Plant InsP3 Levels (pmol/g Fresh Weight)a Stomatal Density (×20 Magnification)b
Wild type 233 ± 3.7 16.4 ± 0.8
Tr 14 ± 1.2 16.7 ± 0.5
a

InsP3 levels were measured in wild-type and transgenic (Tr = combined averages from both T6 and T8) epidermal peels. Data presented are averages ± se of three experiments.

b

Stomatal density was measured in wild-type and Tr epidermal peels at ×20 magnification. Data presented are averages ± se of three experiments.

Stomatal aperture was monitored immediately after peeling or after incubation under dark or light conditions for 3 h (Figure 2A). There was no difference in stomatal aperture between wild-type and transgenic plants under these conditions. We also monitored the response of guard cells to high Ca2+ and external ABA. The transgenic guard cells were less responsive to closing in the presence of 1 mM Ca2+ compared with the wild type (Figure 2B). For the response to ABA, we measured both the promotion of stomatal closure, where epidermal peels were first incubated in the light and then treated with ABA (Figure 2C), and the inhibition of stomatal opening, where epidermal peels were prepared from plants incubated in the dark and then treated with light or ABA in the presence of light (Figure 2D). As seen in Figure 2C, at two different concentrations of ABA (10 and 100 μM), we observed that the transgenic guard cells were more responsive to ABA in the promotion of stomatal closure. However, ABA in the presence of light failed to inhibit stomatal opening in the transgenic guard cells, although opening of wild-type guard cells was strongly inhibited (Figure 2D). These results show that the transgenic guard cells have altered responses to ABA and Ca2+ compared with the wild type and suggest that the differential water loss in the transgenic leaves could in part be regulated at the level of guard cell dynamics (i.e., the higher sensitivity of the transgenic guard cells to ABA-induced closure).

Figure 2.

Figure 2.

The InsP 5-ptase Plants Show Altered Responses to Ca2+ and ABA in Stomatal Bioassays.

(A) Stomatal aperture (presented as the ratio of width to length) was measured in epidermal peels prepared from the abaxial surface of wild-type and transgenic (T8) leaves. Peels were observed directly or after incubation for 3 h in the light or dark.

(B) Effect of high Ca2+. Peels were incubated in buffer with or without Ca2+ or in the presence of 1 mM Ca2+. The asterisk denotes a statistically significant difference compared with the wild type (P < 0.005).

(C) and (D) Effect of ABA. To determine differences in ABA-mediated stomatal closure (C), peels were first incubated in the light for 3 h, and then ABA (10 or 100 μM) was added and the peels were incubated in the light for an additional 3 h. Data are shown as differences between the width:length ratios of the final aperture after ABA treatment and the initial light value (* P < 0.1, ** P < 0.01). To determine differences in inhibition of stomatal opening by ABA (D), peels were taken from plants in the dark (D) and incubated in the light (L) for 3 h in buffer with or without 25 μM ABA. For all stomatal bioassays, ∼15 images were captured per peel (n = 45 per experiment). Data presented are averages ± se of three independent experiments.

Whole Plant Responses to ABA

It has been shown that treatment with ABA can cause an increase in InsP3 in guard cells (Lee et al., 1996) and seedlings (Sanchez and Chua, 2001; Burnette et al., 2003). Similarly, we found that 1- to 2-week-old wild-type seedlings sprayed with 100 μM ABA showed a biphasic increase in InsP3 (Figure 3A, solid black line). InsP3 levels increased rapidly with spraying and again after ∼30 min. Plants were sprayed with water containing 0.1% ethanol as the solvent control (Figure 3A, dotted black line). InsP3 levels were significantly reduced in the transgenic plants and showed no appreciable increase over the time course after spraying with ABA or solvent control (Figure 3A, solid gray and dotted lines).

Figure 3.

Figure 3.

InsP3 and Gene Expression Changes in Response to Exogenous ABA.

(A) Time course of InsP3 changes in response to ABA. Two-week old Arabidopsis seedlings were sprayed with either water containing 0.1% ethanol as the solvent control (dashed line) or 100 μM ABA (solid line) and harvested at the indicated times for InsP3 analysis. Data presented are averages ± se of four independent experiments assayed in duplicate. The changes in InsP3 are plotted as percentages of the wild-type basal level.

(B) Representative data of gene expression changes in response to ABA. RNA was isolated from 2-week-old seedlings sprayed either with water or ABA, and RT-PCR was performed using primers specific for RAB18, KIN2, and COR15a. The Arabidopsis polyubiquitin gene (UBQ10) was used as a control. Similar results were obtained in two independent experiments.

(C) Inhibition of seed germination by ABA. Seed germination assays were performed with increasing concentrations of ABA. Germination was scored after 3 d, and the data presented are averages ± se of three independent experiments.

We also monitored the expression of ABA-responsive genes after treatment with ABA and detected no impairment in the induction of select ABA-responsive genes in the transgenic lines compared with the wild type (Figure 3B). Expression of several ABA-responsive genes, including RAB18, KIN2, and COR15a (Kawaguchi et al., 2004; Shinozaki and Yamaguchi-Shinozaki, 2007), was induced equally well (if not better), with no delay in timing in the transgenic plants compared with the wild type. These results indicate that an ABA-mediated increase in InsP3 is not essential for the induction of these ABA-responsive genes. We also monitored the seed germination response of transgenic lines compared with the wild type to increasing concentrations of ABA. Seed germination in the transgenic lines was less sensitive to 5 and 10 μM ABA compared with the wild type (Figure 3C); however, at 25 μM ABA, germination was dramatically inhibited in both transgenic and wild-type plants.

Changes in Inositol Phosphate Metabolism

In addition to InsP3, it has been reported that inositol hexakisphosphate (InsP6) is a signaling molecule in guard cells and that InsP6 levels increase rapidly following ABA treatment (Lemtiri-Chlieh et al., 2000, 2003). Enzymes involved in the conversion of InsP3 to InsP6 have been functionally characterized in Arabidopsis (Stevenson-Paulik et al., 2002; Xia et al., 2003). Therefore, we set out to determine the impact of increased InsP3 hydrolysis on the biosynthesis of inositol phosphate and, in particular, InsP6. Arabidopsis seedlings were labeled with [3H]inositol for 3 to 4 d, and the soluble inositol phosphates were extracted and analyzed by anion-exchange HPLC. The [3H]inositol labeling results from whole seedlings (Figure 4A) or epidermal peels (Figure 4B) show altered levels of [3H]inositol phosphates in the transgenic lines compared with the wild type. In the transgenic seedlings, consistent with increased InsP3 hydrolysis, InsP2 levels were slightly elevated and total InsP3 isomer levels were decreased. Total InsP5 isomer levels were also decreased. Under these labeling conditions, [3H]InsP4 species were undetectable over background in both wild-type and transgenic plants and thus are not shown in the figure. By contrast, InsP6 levels were readily detectable and were reduced in the transgenic seedlings (by 25%) and peels (by 50%) compared with the wild type. These results support a model for a direct route of InsP6 biosynthesis from InsP3 via the inositol polyphosphate kinases (Stevenson-Paulik et al., 2002; Raboy, 2003). The decrease in InsP6 in the epidermal peels of the transgenic plants could contribute to the altered ABA and Ca2+ responses of the guard cells, since InsP6 has been shown to affect stomatal regulation (Lemtiri-Chlieh et al., 2000, 2003).

Figure 4.

Figure 4.

InsP6 Is Reduced in InsP 5-ptase Seedlings and Epidermal Peels Compared with Wild-Type Plants.

(A) Inositol phosphate profiles from 1-week-old seedlings (wild type and T8) labeled for 4 d and analyzed by anion-exchange chromatography. Data presented are averages ± se of five independent experiments (* P < 0.1).

(B) Inositol phosphate profiles from epidermal peel fragments labeled for 18 h. Data presented are averages ± se of three independent experiments (* P < 0.1, ** P < 0.01).

The Onset of Water Stress Is Delayed in the Transgenic Plants

To further characterize the drought response, we monitored the relative water content (RWC) of wild-type and transgenic leaves during a period of dehydration (Griffiths and Bray, 1996). For these experiments, well-watered plants of similar size and maturity (∼6 weeks old) were maintained in a growth chamber under short-day conditions at a relative humidity of 30%. RWC was measured on day 0 (well watered) and over a period of 8 d without water. A RWC of <60% is considered to be a severe water stress (Kawaguchi et al., 2003). As seen in Figure 5A, the transgenic lines maintain a RWC of >85% for longer than the wild type.

Figure 5.

Figure 5.

The InsP 5-ptase Plants Maintain Higher Relative Water Contents and Lower ABA Levels Compared with the Wild Type during Drought Stress.

(A) The RWC of leaves was measured on day 0 (well watered) and over a period of 8 d without water. Each time point represents the average value of 10 separate plants per line, and the data presented are averages ± se from three independent experiments.

(B) ABA levels were measured in samples from day 0, 7, and 8 from three independent experiments. Data presented are averages ± se from both the wild type and vector control (WT/C2) and from the two transgenic lines T6 and T8 (Tr).

We also monitored ABA levels in the plants during the RWC time course. ABA levels were measured in leaf samples harvested at day 0 and at days 7 and 8 (Figure 5B). At day 0, ABA levels were equally low in all plant lines (see Supplemental Figure 3 online). As the drought stress progressed and the RWC decreased, the level of ABA increased in both wild-type and transgenic lines. However, ABA levels always remained lower in the transgenic plants compared with the wild type even at a RWC of ∼60%. The higher RWC and lower ABA levels at both days 7 and 8 of a no-watering regime (Figure 5) indicate that the transgenic plants are experiencing less drought stress than the wild-type controls.

In some trials, the plants were rewatered on day 8. Upon rewatering, the transgenic plants recovered fully compared with the wild-type plants (see Supplemental Figure 4A online). The faster recovery is also evident by the expression patterns of stress-inducible genes such as RAB18 and ERD1 (see Supplemental Figure 4B online). Both wild-type and transgenic plants showed high expression of these genes at day 8 (when the RWC was ≤60%). However, upon rewatering, the expression of these genes decreased in the transgenic plants while remaining elevated in the wild type.

Microarray Analysis of Transcript Changes with Drought Stress

In order to compare global expression profiles between wild-type and transgenic plants in response to water stress, we harvested leaves for transcript profiling at day 0 (well watered) and at day 7, when the transgenic plants were still at a RWC of >85% and the wild-type and vector control plants were at a RWC of ∼50 to 60%. For each biological experiment, leaves were harvested and pooled from ∼10 plants of each line/time point. Three biological replicates were performed with the wild type and two independent transgenic lines (T6 and T8). Two biological replicates of the vector control line (C2) were performed as an additional control.

Transcript profiling was performed using the Arabidopsis ATH1 arrays (Affymetrix). Array hybridization, data acquisition, and analysis were performed by Expression Analysis of Durham, NC, using the Affymetrix fluidics station and GCOS software. As indicated in Figure 6A, two-group comparisons were made in order to determine any differences in expression profiles between the basal (0) and drought-stressed (D) samples for each line and also between the basal expression profiles of the wild-type and transgenic lines. The two-group comparison incorporates a permutation analysis for differential expression and an estimate of the false discovery rate (Pawitan et al., 2005; Clarke and Zhu, 2006).

Figure 6.

Figure 6.

Schematic of the Microarray Experiment and Confirmation of Select Transcript Profiles Using qRT-PCR.

(A) Schematic showing the control (0) and drought-stressed (D) samples analyzed by microarray and the comparisons (numbered arrows) that were performed between samples.

(B) to (D) qRT-PCR of select transcripts to validate the array results.

(B) Select transcripts showing basal downregulation in transgenic lines compared with the wild type and vector control.

(C) Select transcripts showing basal upregulation in transgenic plants compared with the wild type and vector control.

(D) Select transcripts that are upregulated by drought stress in the wild type and vector control are significantly greater compared with transgenic plants.

The qRT-PCR results for (B) and (C) are presented as fold change compared with the wild type and are averages ± se from four independent biological samples. For (D), data are presented as percentage fold change of the wild type and are averages ± se of three independent biological replicates (wild-type drought value was set to 100%).

Comparison of Basal Expression Profiles

A comparison of the basal expression profiles of the two transgenic lines [T8(0) versus T6(0), comparison 7, Figure 6A] did not reveal any significant differences (false discovery rate > 0.9), which indicates that the transgenic lines are comparable and similar, thereby ruling out potential positional effects. We next looked at differences in basal expression profiles between transgenic and wild-type plants. Transcripts that were commonly upregulated or downregulated in both transgenic lines compared with the wild type (based on comparisons 4 and 5, Figure 6A) were further checked against basal levels of expression in the vector control line (C2) to ensure that the differences were specific for the transgene and not due to transformation.

The transgenic plants show no obvious morphological or developmental differences compared with wild-type plants under normal growth conditions. Consistent with this result, only a small fraction of transcripts showed altered basal expression in the transgenic lines compared with the wild type and the vector control (<0.4% of the transcriptome). There were more downregulated genes (62 transcripts; see Supplemental Table 1 online) compared with upregulated genes (20 transcripts; see Supplemental Table 2 online) in the transgenic lines. Of the downregulated genes, 18% fall into the defense-related category, including several pathogenesis-related genes (PR-1, PR-2, and PR-5). A second distinct group of downregulated genes (11%) represent proteins involved in Ca2+ binding and transport and protein folding, including CRT3 and BiP3 (Table 2). This downregulation may be an indication that the InsP3-sensitive Ca2+ stores are underutilized in the transgenic plants due to the rapid metabolism of InsP3. Conversely, in tobacco cells expressing the highly active human PtdInsP kinase Iα (which have basal InsP3 levels ∼ 50 fold higher than the wild type), we detected an upregulation of the endoplasmic reticulum Ca2+ binding proteins, presumably to maintain a high-capacity Ca2+ store (Im et al., 2007). We confirmed the transcript profile results for select genes from each category using quantitative RT-PCR (qRT-PCR), as shown in Figure 6B.

Table 2.

Subset of Genes Involved in Ca2+ Binding and Endoplasmic Reticulum Protein Folding and Transport That Are Basally Downregulated in the Transgenic Plants

Locus Identifier Fold Change P Predicted Function and Gene Name
At2g18660 −9.15 0.01710 Expansin family protein (EXPR3)
At3g47480 −4.12 0.02440 Calcium binding EF hand family protein
At1g09080 −3.87 0.00640 Luminal binding protein 3 (BiP-3)
At3g60540 −3.57 0.00710 sec61 β family protein
At1g77510 −3.09 0.00350 Protein disulfide isomerase, putative
At4g16660 −2.57 0.01040 Heat shock protein 70, putative/HSP70, putative
At4g24920 −2.43 0.01090 Protein transport protein SEC61 γ subunit, putative
At1g08450 −2.40 0.00530 Calreticulin 3 (CRT3)
At3g51860 −2.35 0.00090 Cation exchanger, putative (CAX3)
At4g24190 −2.16 0.01730 Shepherd protein (SHD)/clavata formation protein, putative
At5g50460 −2.07 0.00900 Protein transport protein SEC61 γ subunit, putative

Subset of the basally downregulated genes in the two InsP 5-ptase transgenic lines (T6 and T8) compared with wild-type and C2 plants revealed by the microarray analysis. Data presented are from transgenic line T8. Transcripts were selected based on a greater than twofold decrease (P < 0.1).

One of our more remarkable and unexpected findings is that the drought-responsive transcription factor DREB2A and a subset of DREB2A-regulated genes are basally upregulated in the transgenic plants. Of the 20 genes showing twofold higher basal expression in the InsP 5-ptase plants compared with wild-type and vector control plants (see Supplemental Table 2 online), the genes shown in Table 3 were found to be upregulated by greater than ninefold in Arabidopsis plants expressing a constitutively active form of DREB2A (Sakuma et al., 2006b). The basal upregulation in the InsP 5-ptase transgenic plants of the six genes shown in Table 3 was confirmed using qRT-PCR and additional independent biological replicates (Figure 6C).

Table 3.

Subset of DREB2A-Regulated Genes That Are Basally Upregulated in the Transgenic Plants

Locus Identifier Fold Change P Predicted Function and Gene Name
At5g05410 3.20 0.0001 DRE binding protein (DREB2A)
At3g55940 2.47 0.0012 Phosphoinositide-specific phospholipase C, putative (PLC7)
At4g36010 2.46 0.0238 Pathogenesis-related thaumatin family protein
At1g09350 2.36 0.0231 Galactinol synthase, putative (GolS3)
At1g56600 2.24 0.0857 Galactinol synthase, putative (GolS2)
At3g50970 2.05 0.0756 Dehydrin xero2 (XERO2)/low temperature–induced protein (LTI30)

Subset of the basally upregulated genes in the two InsP 5-ptase transgenic lines (T6 and T8) compared with wild-type and C2 plants revealed by the microarray analysis. Data presented are from transgenic line T8. Transcripts were selected based on a greater than twofold increase (P < 0.1).

Comparison of Expression Profiles in Response to Drought Stress

The two-group comparison analysis between the basal (0) and drought-stressed (D) samples (comparisons 1, 2, 3, and 6) were used to compare gene expression profiles in response to drought stress. We generated four-way Venn diagrams using the upregulated or downregulated lists for each plant line and determined the common and unique patterns of expression between wild-type and transgenic lines in response to drought stress. The lists generated from the Venn comparison were further checked for overlap between the wild-type/C2 only and transgenic only groups (for genes that showed similar greater than twofold upregulation or downregulation, although with higher P values) to generate the lists of upregulated or downregulated genes unique to either the wild-type and vector control only or transgenic lines only (see Supplemental Tables 3 to 6 online). Table 4 summarizes the numbers of genes that fall into each of these categories. Interestingly, there was a greater number of upregulated or downregulated genes in the wild-type/C2 group than in the transgenic lines, which could be a reflection of the level of perceived stress (Table 4). For both upregulation and downregulation, there were ∼150 to 200 genes that were either upregulated or downregulated with drought in all four plant lines tested (the complete lists of commonly upregulated and downregulated genes and pie charts depicting their functional categorization based on Gene Ontology annotation are shown in Supplemental Tables 7 and 8 and Supplemental Figures 5A and 5B online).

Table 4.

Transcripts That Are Upregulated or Downregulated with Drought in Wild-Type and Transgenic Plants

Regulation Wild Type/C2 Only Common Tr Only
Upregulated with drought greater than twofold 244 219 4
Downregulated with drought greater than twofold 183 161 22

There were 244 transcripts that were exclusively upregulated (greater than twofold) with drought stress in both wild-type and vector control plants (see Supplemental Table 3 online). Interestingly, 30% of these transcripts are known to be regulated by ABA (Li et al., 2006b) and include a subset of late embyrogenic abundant (LEA) transcription factors and PP2C protein phosphatases. Four of five of this subset of LEAs belong to the same group recently classified as the LEA_4 group (Hundertmark and Hincha, 2008). The array results for some of the transcripts in Supplemental Table 3 online that are also regulated by ABA were confirmed by qRT-PCR (Figure 6D). By contrast, there were only four genes that were uniquely upregulated in the transgenic lines with drought (see Supplemental Table 4 online). The significance of this set of genes and what they share in common is not known at this point. In the downregulated with drought category, wild-type/C2 uniquely had 183 genes (listed in Supplemental Table 5 online). In the transgenic lines, there were 22 uniquely downregulated genes with drought (see Supplemental Table 6 online). These are predicted to encode unknown proteins (25%), proteins involved in carbohydrate metabolism (17%), transport-related proteins (16%), signaling proteins (14%), transcription factors (11%), protein modification (10%), and calmodulin-related proteins (6%). The signaling-related proteins include a mitogen-activated protein kinase (At MPK11) that was reduced by approximately threefold in the transgenic plants compared with the wild type.

InsP3-Mediated Ca2+ Signaling Is Attenuated in the Transgenic Plants

A well-characterized downstream consequence of increased InsP3 is the release of Ca2+ from intracellular stores. Transient and rapid increases in intracellular Ca2+ have been demonstrated in response to many abiotic stresses, including cold, salt, and osmotic stress (Knight et al., 1996, 1997a; Knight and Knight, 2001). We predicted that if InsP3 is involved in mediating a stress-induced Ca2+ signal, then the transgenic plants would have impaired Ca2+ signaling in response to the stress. In order to test this hypothesis, we generated transgenic plants expressing the Ca2+ binding photoprotein aequorin (Knight et al., 1997b). Two independent aequorin-expressing control lines and two independent aequorin-expressing InsP 5-ptase lines were selected and tested for Ca2+ response. Seedlings were incubated in the substrate coelenterazine and then subjected to a salt or a cold stimulus (by injecting 250 mM NaCl or ice-cold water), and luminescence was measured in a luminometer. Luminescence counts were collected every 0.2 s and converted to Ca2+ concentration as described (Knight et al., 1997b). The peak Ca2+ response was reduced in the InsP 5-ptase lines compared with the control in response to a cold or a salt stimulus (Figure 7). The Ca2+ response measured in these experiments is a reflection of the stress-induced increase in intracellular Ca2+ as result of entry from the outside as well as release from intracellular stores. Since InsP3 triggers the release of Ca2+ from intracellular stores, these experiments reveal the InsP3-mediated component of Ca2+ release from intracellular stores. Based on our results, the InsP3-mediated component of the Ca2+ signal in these plants is ∼150 to 200 nM (Table 5) and constitutes ∼30% of the total Ca2+ signal under these conditions.

Figure 7.

Figure 7.

The InsP 5-ptase Plants Show an Attenuated Ca2+ Response to a Salt or a Cold Stimulus.

(A) Aequorin-expressing control (wild-type) and InsP 5-ptase (Tr) seedlings were treated with 0.25 M NaCl, and luminescence was measured in a luminometer.

(B) A similar experiment in which seedlings were treated with ice-cold water.

Representative data of salt- and cold-induced Ca2+ release are shown. Each trace is the average from three to four seedlings. The solid lines are traces from the stimulus, and the dotted lines are traces from adding water, which controls for mechanical or touch stimulation. The experiments were repeated four to five times with similar results.

Table 5.

Peak Ca2+ Response to a Cold, Salt, or ABA Stimulus

Measurement Colda
Saltb
ABAc
Wild Type Tr Wild Type Tr Wild Type Tr
Peak Ca2+ concentration (μM) 0.60 ± 0.1 0.35 ± 0.04 0.65 ± 0.04 0.48 ± 0.01 0.30 ± 0.02 0.22 ± 0.01
n 10 12 8 8 15 15

The average peak for the water control for the cold and salt experiments was 0.17 ± 0.02 μM, and the average peak for the solvent control for the ABA experiments (0.1% ethanol) was 0.11 ± 0.01 μM.

a

Average ± se of 10 to 12 experiments using two control (wild type) and two InsP 5-ptase (Tr) lines.

b

Average ± se of eight experiments using two control (wild type) and two InsP 5-ptase (Tr) lines.

c

Average ± se of 15 experiments using two control (wild type) and two InsP 5-ptase (Tr) lines.

In separate experiments, aequorin-expressing wild-type and transgenic seedlings were treated with 100 μM ABA. Although the whole seedling Ca2+ response to ABA was less dramatic than with the salt or cold stimulus, there was a similar ∼30% decrease in the peak Ca2+ in the transgenic seedlings compared with the control (Table 5).

DISCUSSION

In order to examine the downstream consequences of a constitutive reduction in InsP3, we generated transgenic Arabidopsis plants expressing the mammalian type I inositol phosphatase. This enzyme is specific for InsP3 hydrolysis and is involved in terminating the InsP3 signal in animal cells (Mitchell et al., 1996; De Smedt et al., 1997). In the transgenic plants, basal levels of InsP3 were ∼5% of wild-type values (Perera et al., 2006) and InsP3 remained below basal wild-type values even in response to stress.

InsP3-Mediated Ca2+ Signaling Is Attenuated in the InsP 5-ptase Plants

According to the signaling paradigm, the first documented downstream effect of InsP3 is to trigger the release of Ca2+ from intracellular stores. Using aequorin-expressing seedlings, we demonstrated that the InsP3-mediated Ca2+ signal is reduced in the transgenic plants compared with wild-type plants in response to salt, cold, or exogenously added ABA and that the InsP3-mediated Ca2+ contribution is ∼30%. Aequorin-based technology is best suited for monitoring changes in Ca2+ with fast kinetics on the order of seconds to minutes; therefore, it was not possible to monitor gradual changes that may occur over a period of days with a physiological drought stress. Nevertheless, it is clear from our results that Ca2+ signaling and homeostasis are subtly altered in the transgenic plants. The transgenic guard cells were less responsive to high Ca2+, and there was a downregulation of genes involved in Ca2+ binding and transport.

ABA Responses Are Altered in the Transgenic InsP 5-ptase Plants

It is well established that Ca2+ is a key intermediary in some but not all of the ABA-regulated processes. ABA is known to inhibit K+-inward rectifying channels, a process that also involves an increase in Ca2+ and inositol phosphates (Ng et al., 2001; Hetherington and Brownlee, 2004). The pH-induced anion efflux in guard cells, however, is independent of Ca2+, and activation of guard cell anion efflux channels by the application of ABA to the cytosol was not preceded by an elevation of Ca2+ and could not be mimicked by inositol phosphates or cADP-ribose (Levchenko et al., 2005).

The transgenic InsP 5-ptase guard cells showed an impaired response to the ABA inhibition of stomatal opening compared with wild-type plants. The decreased response of the transgenic guard cells to ABA and to high Ca2+ may also be explained by the fact that InsP6 is reduced, since InsP6 is important for guard cell regulation (Lemtiri-Chlieh et al., 2000, 2003). Plants with reduced PLC were similarly insensitive to ABA in the stomatal opening assay (Mills et al., 2004). Taken together, these results support the speculation that PLC (and/or InsP3 and InsP6) may only be required for the ABA-mediated inhibition of the opening response (Mills et al., 2004).

By contrast, the stomatal closure response to exogenous ABA was enhanced in the InsP 5-ptase plants. This could enhance sensitivity in the drought response, even though in planta the endogenous ABA levels did not increase as rapidly with drought stress. It is clear that guard cell regulation is a complex process involving several partially or completely independent pathways (Schroeder et al., 2001; Roelfsema and Hedrich, 2005; Israelsson et al., 2006; Li et al., 2006a) and that ABA-mediated opening and closure, although related, may be regulated via distinct mechanisms (Christmann et al., 2006; Li et al., 2006a).

With regard to ABA-mediated signaling, the induction of ABA-responsive genes with exogenous ABA treatment was comparable (if not stronger) in the InsP 5-ptase transgenic and wild-type plants. This is in contrast with plants overexpressing At5PTase1 and At5PTase2 (Sanchez and Chua, 2001; Burnette et al., 2003), which showed delayed or attenuated expression of select ABA-responsive genes.

We suspect that the unique features of the mammalian type I InsP 5-ptase enzyme used in our studies may help explain the differential responses observed with the InsP 5-ptase plants compared with the plant counterparts. Both At5PTase1 and At5PTase2 have been shown to hydrolyze the membrane phospholipids in addition to InsP3 (Ercetin and Gillaspy, 2004), while the animal enzyme is specific for the soluble inositol phosphates such as InsP3 (Majerus et al., 1999). Changes in stomatal aperture are regulated by volume changes of the guard cell, which could involve up to a twofold change in membrane surface area (Kubitscheck et al., 2000). This, in turn, would require significant reorganization of membrane lipids via vesicle secretion and endocytosis (Schroeder et al., 2001). Overexpressing At5PTases or decreasing production of PLC could have a negative impact on membrane biogenesis and, thereby, stomatal regulation. Guard cell responses are affected by PtdIns3P and PtdIns4P (Jung et al., 2002), and PtdInsP2 was recently shown to be important for regulating stomatal opening (Lee et al., 2007). The reduced stomatal opening observed in the At5TPase1-overexpressing plants could in part be due to reduced levels of PtdInsP2, although Arabidopsis 5ptase1 null mutants do not show any increases in PtdInsP2 levels (Gunesekera et al., 2007). The effect of increased PtdInsP2 was evident in the sac9 mutant (which is defective in a PtdInsP2-hydrolyzing 5-phosphatase and overaccumulates PtdInsP2), which shows a constitutive stress phenotype with closed stomata (Williams et al., 2005). PtdInsP2 homeostasis is clearly important for normal growth and development in plants, and altering this balance can lead to severe consequences.

When comparing other methods for altering endogenous InsP3, it is important to note that the inhibition of PLC will have far-reaching effects on phosphoinositide turnover in addition to reducing InsP3, making it difficult to compare our results with those from studies in which InsP3 is lowered by blocking PLC. Reduced PLC hydrolysis could lead to a buildup of PtdInsP2 (unless the lipid phosphatases are activated) and would also have an impact on phosphatidic acid, either via diacylglycerol or by activating specific phospholipase D isoforms (Qin and Wang, 2002). Phosphatidic acid plays an important role in promoting ABA-mediated guard cell closure by the interaction with both ABI1, a negative regulator of ABA (Zhang et al., 2004, 2005), and Gpa1 (Mishra et al., 2006). Phosphatidic acid levels have been shown to increase in response to hyperosmotic stress (Munnik et al., 2000) as well as several other stresses (Wang, 2004; Testerink and Munnik, 2005), and blocking this increase may adversely affect the plant's ability to respond to stress.

Basal and Drought-Induced Transcriptional Changes in the InsP 5-ptase Transgenic Plants May Contribute to the Enhanced Drought Tolerance

The microarray results have revealed interesting and important differences in transcript abundance between the wild-type and transgenic plants both in basal levels and also in the response to drought stress. The most striking result in the basal comparison is the upregulation of DREB2A and the subset of DREB2A-regulated genes in the transgenic plants. DREB2A belongs to a family of AP2 domain–containing transcription factors. The DREB1/CBF group is involved in cold stress (Thomashow, 1999), while DREB2A and DREB2B have been implicated in responses to dehydration, salt, and, more recently, heat stress (Sakuma et al., 2006a, 2006b; Schramm et al., 2007). In early experiments (Liu et al., 1998), DREB2A expression was found to be highly induced by dehydration and salt treatment, but not by ABA, leading to the conclusion that DREB2A functions mainly in an ABA-independent drought stress–responsive pathway (Shinozaki and Yamaguchi-Shinozaki, 2000, 2007; Yamaguchi-Shinozaki and Shinozaki, 2006). DREB2A has been shown to bind to DRE elements in stress-responsive gene promoters; binding is regulated by a negative regulatory domain in the DREB2A protein (Sakuma et al., 2006b) and may also be influenced by phosphorylation (Agarwal et al., 2007). Although constitutive overexpression of specific DREB family transcription factors conferred improved stress tolerance, it caused severely retarded growth under normal conditions (Liu et al., 1998; Kasuga et al., 1999). However, if only a subset of these genes were upregulated, it might be possible to induce tolerance without impairing growth. We propose that the constitutive dampening of the InsP3 signal and the increased turnover of pathway intermediates could lead to the selective derepression of stress-induced transcripts and to increased stress tolerance without inhibiting growth under nonstressed conditions.

In the InsP 5-ptase plants, there were five basally upregulated genes (excluding DREB2A), and all of these contain a DRE binding motif in their promoter sequences (Sakuma et al., 2006b). Three of the genes (XERO2, GolS2, and GolS3) are implicated in osmoprotection and drought tolerance. Xero2 is a dehydrin, and galactinol synthase catalyzes the first committed step in the synthesis of the raffinose family of oligosaccharides (RFOs). RFOs are known to protect seeds during desiccation (Obendorf, 1997). In Arabidopsis, the At GolS1, -2, and -3 genes are upregulated by drought and temperature stress (Taji et al., 2002; Nishizawa et al., 2006), and plants overexpressing At GolS2 showed reduced transpiration, increased drought tolerance, and an increase in galactinol and raffinose sugars in leaves (Taji et al., 2002). Another factor that is critical for the production of RFOs is the availability of myo-inositol. Mutant soybean (Glycine max) having reduced myo-inositol had drastically reduced galactinol and RFOs (Hitz et al., 2002; Karner et al., 2004). It is possible that the increased hydrolysis of InsP3 in the transgenic InsP 5-ptase plants could lead to an increased flux through this pathway, resulting in increased levels of RFOs. Preliminary work suggests that basal levels of raffinose and inositol are similar in InsP 5-ptase plants compared with wild-type plants (see Supplemental Table 9 online).

It is curious that a putative PLC gene (At PLC7) is upregulated, although this gene is not predicted to encode a functional protein (Hunt et al., 2004); therefore, the significance of this result is unclear at present. In addition to the six listed genes, two other DREB2A-regulated genes, COR47 (At1g20440) and LEA14 (At1g01470), were found to be expressed by >1.5 fold in the transgenic lines over the wild type and vector control (P < 0.05). Basal upregulation of select stress-responsive genes was found to be one of the critical differences between Arabidopsis and its halotolerant relative, Thellunigiella halophila (Taji et al., 2004; Gong et al., 2005). It is likely that the basal upregulation of select DREB2A-regulated genes in the transgenic InsP 5-ptase plants contributes to their increased ability to withstand drought stress.

Comparison of the drought-stressed samples has also revealed some important insights. There were ∼150 to 200 genes that were commonly upregulated or downregulated in both transgenic and wild-type plants. Many of these are consistent with other published studies and show similar patterns of expression with abiotic stress. The list of commonly upregulated transcripts from this study shares 75% overlap with a recently reported list of drought-induced genes in wild-type plants (Catala et al., 2007). These genes probably represent pathways and processes that are not altered in the InsP 5-ptase plants. In a comparison of three separate microarray studies on osmotic and drought stress, Bray (2004) found that there were only 27 commonly upregulated genes. Since the experimental methods and developmental stages of plants used varied between the experiments (Seki et al., 2001; Kreps et al., 2002; Kawaguchi et al., 2004), it is not surprising that this number is low. Interestingly, in this work, we found 15 of the common 27 genes upregulated in both the transgenic and wild-type/C2 plants, and 5 of these 15 transcripts showed a >10 fold increase in all plant lines with drought stress.

Further support for ABA-independent pathways contributing to drought tolerance in the transgenic plants comes from the observation that, of the genes that were uniquely upregulated in wild-type/C2, ∼30% are known to be regulated by ABA (Li et al., 2006b) and many of these did not show any increase in the transgenic plants with drought stress. It is possible that the expression of these genes requires a certain threshold of ABA that has not been reached in the transgenic plants, or, alternatively, that induction of these specific genes is impaired in the transgenic plants. There were four genes that were uniquely upregulated by drought stress in the transgenic plants, and it will be interesting to determine if these genes have any common features and if their patterns of expression are similar in the transgenic plants when the drought stress is prolonged.

In related work, the constitutive expression of the InsP 5-ptase in tomato (Solanum lycopersicum) plants resulted in plants with increased drought tolerance (M. Khodakovskaya, C. Sword, I.Y. Perera, W.F. Boss, C.S. Brown, and H. Winter-Sederoff, unpublished data). In contrast with the Arabidopsis plants, the transgenic tomato plants have increased biomass and increased levels of soluble sugars, suggesting that basal metabolism is differentially sensitive to the decrease in InsP3 in tomato versus Arabidopsis. The increase in soluble sugars seen in the InsP 5-ptase tomato plants may play a protective role during drought stress. Stress tolerance is often attributed to the ability to maintain higher levels of several metabolites, including hexose sugars, in both the presence and absence of stress (Bartels and Sunkar, 2005; Valliyodan and Nguyen, 2006). Because we did not detect comparable differences in soluble sugars in the InsP 5-ptase Arabidopsis plants under normal conditions (see Supplemental Table 9 online), we anticipate that the mechanisms of tolerance in the two plants are different. Further comparisons will be necessary to identify the InsP3-mediated response in these two systems.

In summary, although constitutively decreasing InsP3 was predicted to render plants more sensitive to drought stress, our findings indicate that decreasing InsP3-mediated signaling leads to increased drought tolerance. The increased drought tolerance in the transgenic plants cannot be explained by the altered responses to ABA. Apart from the increased response to exogenously added ABA in the stomatal closure assay, the transgenic plants show reduced ABA levels with drought stress and lack of induction of several ABA-regulated genes compared with the wild type. However, the transgenic plants show a compensatory upregulation of an ABA-independent pathway involving DREB2A and a select subset of DREB2A-regulated genes. These results suggest that the coordinated regulation of these genes in the InsP 5-ptase plants contributes to conferring drought tolerance without adversely affecting plant growth. The selective regulation of this subset of the DREB2A pathway could be promising as a mechanism for increasing drought tolerance in crop plants.

METHODS

Plant Material and Growth Conditions

Wild-type (ecotype Columbia) and InsP 5-ptase transgenic Arabidopsis thaliana plants (Perera et al., 2006) were grown in the North Carolina State University Phytotron in a growth chamber under short-day conditions (8 h of light/16 h of dark) at 21°C with a light intensity of ∼150 μmol·m−2·s−1. Six-week-old, fully-grown plants (prior to bolting) were used for most experiments. For all soil-grown experiments, a large batch of soil mix (Promix PGX; Hummert International) was moistened well with water and the pots were filled with an equal amount of soil prior to planting the seeds. For experiments using seedlings, seeds were surface-sterilized as described previously (Perera et al., 2006) and stratified for 48 h at 4°C prior to plating on Murashige and Skoog medium (Caisson Labs) containing 1% sucrose and 0.8% agar type M (Sigma-Aldrich). Plates were incubated vertically in a growth chamber under short-day conditions as described above.

Measurements of Water Loss and Drought Stress

Leaf water loss was measured in fully expanded leaves of comparable size, weight, and development that were excised from the rosettes of well-hydrated wild-type and transgenic plants. The excised leaves (adaxial side up) were kept at 25°C room temperature in the light (30% humidity) for 1 to 2 h. Fresh weight of the leaves was monitored at 5-min intervals. After the water loss measurements, the surface area of the leaves was measured using a LI 3100 area meter (Li-Cor). For whole plant water loss measurements, plants were well watered at the start of the experiment and the surface of the pot was covered with plastic wrap to prevent any evaporation from the soil. Pots were weighed twice daily (in the morning just prior to when the lights came on in the chamber and immediately after the lights were turned off in the evening) for 6 d. Stomatal conductance was measured using an AP4 cycling porometer (Dynamax). At least three to four measurements were made per individual plant, and four to five plants per line were monitored. RWC was measured as described (Griffiths and Bray, 1996) using 6-week-old plants over a period of 8 d without water in a chamber with relative humidity of 30%. (RWC = [FW − DW]/[TW − DW], where FW = fresh weight of leaf, DW = dry weight of leaf, and TW = rehydrated weight of leaf after incubating in water overnight.) The average value for 10 separate plants per line is indicated at each time point.

Stomatal Bioassays

Epidermal peels were prepared from the abaxial surface of wild-type and transgenic Arabidopsis leaves in buffer (5 mM MES/KOH, pH 6.5, 50 mM KCl, and 0.1 mM CaCl2) as described by Roelfsema and Prins (1995). Guard cells were observed using a 100×/1.25 oil objective on a Leica DMLS microscope. Aperture widths and lengths were measured, and stomatal apertures were determined as the ratio of width to length. For each treatment, ∼45 measurements (from three separate peels) were collected, and each experiment was repeated at least three times. Peels were observed directly (after preparing peels from plants kept in the light) or incubated in the light for 3 h. For dark measurements, plants were kept in darkness for 3 h prior to making the epidermal peels. For Ca2+ treatment, peels were incubated for 3 h in the light in the buffer described above (0.1 mM Ca2+) or in this buffer with no added Ca2+ or with 1 mM Ca2+. To determine differences in ABA-mediated stomatal closure, peels were first incubated in the light for 3 h, and then 10 or 100 μM ABA (cis,trans-ABA from A.G. Scientific) was added and the peels were incubated in the light for 3 h. Data are shown as the difference between the width:length ratios of the final aperture after ABA treatment and the initial light value. The differences in inhibition of stomatal opening by ABA were determined using peels taken from plants that had been in the dark for 3 h and that were then incubated in the light for 3 h in buffer with or without 25 μM ABA.

Seed Germination Assays

Surface-sterilized, stratified seeds were plated on Murashige and Skoog plates containing a filter paper soaked in water or different concentrations of ABA as described (Weigel and Glazebrook, 2002). Germination was scored as the emergence of green cotyledons at 3 to 4 d after plating.

ABA Measurements and Sugar Analysis

ABA levels and soluble sugars were measured in leaf samples from the RWC time course. Leaves were frozen and ground in liquid N2 and extracted in a 60:40 methanol:water solution. Samples were analyzed at the Metabolomics and Proteomics Laboratory at North Carolina State University. ABA was measured using a Thermo LTQ ion-trap mass spectrometer operating in electrospray ionization mode. Briefly, a 5-μL aliquot was injected onto a 150- × 2-mm Pursuit XRs column (Varian) equilibrated in 80:20 A:B, where A = 0.1% ammonium formate in water, pH 3.5, and B = acetonitrile, with a flow rate of 250 μL/min. A linear gradient to 65% B was utilized to elute ABA. Quantification was against an external standard (reserpine, 2 μM; Sigma-Aldrich), which was spiked into samples. Data were acquired in tandem mass spectrometry mode, ABA m/z 263 > total ion chromatogram; reserpine m/z 609.3 > total ion chromatogram; ABA negative ion mode, reserpine positive ion mode. Soluble sugars and inositol were analyzed by gas chromatography–mass spectrometry. Leaf tissue was ground in cold solvent, mixed with acetonitrile, and dried under vacuum. The sugars were converted to trimethylsilyl derivatives, and gas chromatography–mass spectrometry was performed using a ThermoTrace GC Ultra gas chromatograph coupled to a Thermo DSQ II mass spectrometer. The mass spectrometer was operated with an electron-impact source in positive mode monitoring m/z 191, 204, 217, 361, and 437. Quantitation was conducted by comparing peak areas obtained for trimethylsilyl derivatives of fructose, glucose, and sucrose in the samples with a series of reference standards analyzed concurrently, and data were processed using Thermo's Xcalibur software. Data presented are averages from three independent biological replicates.

Labeling Studies with [3H]Inositol

One-week-old Arabidopsis seedlings (∼10 seedlings per well) were transferred to a multiwell plate containing 800 μL of 0.5× Murashige and Skoog medium containing 45 μCi of [3H]myo-inositol. Plates were incubated in a growth chamber under long-day conditions with gentle rotation to ensure aeration for 4 to 5 d. After incubation, the seedlings were quickly blotted on tissue and ground in liquid N2. The frozen ground powder was incubated in 0.75 n HCl containing 0.2% phytate (as carrier) on ice for 20 min. After incubation, the samples were centrifuged and the supernatants were analyzed for inositol phosphates by anion-exchange chromatography as described by Stevenson-Paulik et al. (2005). Inositol phosphate species were identified based on the coelution of standard [3H]Ins(1,4)P2, [3H]Ins(1,4,5)P3, [3H]Ins(1,3,4)P3, [3H]Ins(1,3,4,5)P4, [3H]Ins(1,4,5,6)P4, [3H]Ins(1,3,4,6)P4, [3H]Ins(1,3,4,5,6)P5, [3H]Ins(1,2,4,5,6)P5, [3H]Ins(1,2,3,4,6)P5, and [3H]Ins(1,2,3,4,5,6)P6. Standards were either purchased from Perkin-Elmer Life Sciences or generated as described (Stevenson-Paulik et al., 2002, 2005). Because many of the inositol phosphate isomers are in low abundance in 3H-labeled Arabidopsis seedlings and cannot be easily resolved using the strong anion exchange column, peaks for [3H]InsP3 and [3H]InsP5 are reported here without specific isomer identification. It should be noted that [3H]InsP4 was not detectable above background in these samples. Epidermal peel fragments were incubated in guard cell incubation buffer (described above, containing 0.15 M mannitol) and 25 μCi of [3H]myo-inositol for 18 h and processed as described above.

InsP3 Assays

Epidermal peels were prepared as described above and incubated in buffer in the light. After ∼2 h, peels were immediately frozen and ground in liquid N2 and assayed for Ins-1,4,5-trisphosphate content using the receptor binding kit (GE Healthcare) as described previously (Perera et al., 2006). For InsP3 measurements after ABA treatment, seedlings were sprayed with either 100 μM ABA or water containing 0.1% ethanol as a solvent control. Samples were harvested at the indicated times, frozen immediately in liquid N2, and assayed for InsP3 content. The cross-reactivity of the bovine brain extract for other InsP3 stereoisomers or InsP4 or InsP5 isomers is <1.0% and <0.03% for InsP6, according to the manufacturer's specifications (http://www.gehealthcare.com/lifesciences). With plant extracts, >90% of the InsP3 content measured with this assay can be removed by treatment with the type I InsP 5-ptase, indicating that the data reported reflect total cellular Ins-1,4,5-P3 (Perera et al., 2001).

Aequorin Measurements

Wild-type Arabidopsis and InsP 5-ptase plants (lines T6 and T8) were stably transformed with Agrobacterium tumefaciens containing the pGIF2-35SAq plasmid (a kind gift from Marc and Heather Knight). Transgenic plants expressing aequorin were selected by RT-PCR and maintained until the T3 or T4 generation. At least two to three aequorin-expressing control and InsP 5-ptase lines were selected, and extracts were tested for aequorin luminescence by in vitro reconstitution as described (Knight et al., 1997b). For stimulation experiments, aequorin-expressing control and InsP 5-ptase seedlings were first incubated overnight in the luminophore (2.5 μM coelenterazine; Nanolight Technology) in the dark with gentle shaking in order to reconstitute functional aequorin. Individual seedlings were transferred to a cuvette, and luminescence was monitored in a Sirius single-tube luminometer (Berthold Detection Systems). The salt stimulus (final concentration of 0.25 M NaCl) or a water control was injected at ∼10 s, and the total reconstitutable Ca2+ was discharged at ∼70 s by a second injection of 2 M CaCl2 in 20% ethanol. Luminescence was monitored at 0.2-s intervals and converted to Ca2+ concentration as described (Knight et al., 1996, 1997b). Each trace is the average from three seedlings. For the cold stimulus, ice-cold water was injected at ∼10 s followed by the Ca2+ discharge at 70 s. For ABA treatments, ABA (final concentration of 100 μM in 0.1% ethanol) was injected at ∼10 s followed by the Ca2+ discharge at 70 s as described above. The data presented in Table 5 are compiled from two independent aequorin-expressing control lines (wild type) and two independent aequorin-expressing InsP 5-ptase lines (one each of T6 and T8; denoted Tr).

Microarray Analysis

For the microarray analysis, leaf samples (pooled from 10 plants per line) were collected on day 0 and day 7 of the RWC time course and frozen immediately in liquid N2. Three biological replicates were performed for the wild type and two independent transgenic lines (T6 and T8) as well as two biological replicates for the vector control (C2). RNA was isolated using the Plant RNeasy kit (Qiagen), and biotinylated target cRNA was synthesized using the two-cycle target-labeling protocol (Affymetrix). Arabidopsis arrays (ATH1 from Affymetrix) were hybridized, and data acquisition and analysis were performed by Expression Analysis using the Affymetrix fluidics station and GCOS software.

Statistical Analysis of Microarray Data

For the two-group analyses, an estimate of signal for each transcript was calculated based on the Microarray Suite 5.0 algorithm (Affymetrix), and raw fold change for each transcript was calculated by taking the simple ratio of the geometric means of the signal values for each respective group (i.e., 0 versus D). For the wild type, T8, and T6, which were represented by three arrays per group, differential expression was determined using a robust implementation of permutation testing as described at the Expression Analysis website (http://www.expressionanalysis.com/pdf/PADE_technote.pdf).

In brief, a modified t-statistic (Di) was calculated for each transcript when comparing groups, and a difference (Δ) was computed between Di and the average or expected t-statistic ordered values from a reference distribution (D[i]), calculated by computing all possible random permutations of our samples. A list of differentially expressed transcripts was then created by selecting for transcripts with an estimated absolute raw fold change of ≥2. The drought-responsive differential expression lists were further selected based on a false discovery rate of ≤5% (Pawitan et al., 2005; Clarke and Zhu, 2006). For the vector control C2 (which had two biological replicates), we used a cutoff of P < 0.05 (comparison 6). For the basal comparisons between the transgenic lines versus the wild type and vector control, where the differential expression lists were limited in size, we used a cutoff of fold change ≥ 2 and P < 0.1 (Tables 2 and 3; see Supplemental Tables 1 and 2 online). Results for several of the genes showing differential basal expression were confirmed using qRT-PCR as described below. To determine the common patterns of expression between the wild type, the vector control, and the two transgenic lines with drought stress, we generated four-way Venn diagrams using the four-way Venn diagram generator at http://www.pangloss.com/seidel/Protocols/venn4.cgi.

RNA Isolation, RT-PCR, and qRT-PCR Analysis

RNA was isolated from harvested leaves using the plant RNeasy Mini kit (Qiagen) with the on-column RNase-free DNase I treatment. The first-strand cDNA was synthesized from total RNA using either the StrataScript QPCR cDNA synthesis kit (Stratagene) or the High-Capacity cDNA Archive kit (Applied Biosystems) and random primers. For RT-PCR, cDNAs were amplified using HotStar Taq DNA Polymerase (Qiagen) and gene-specific primers (see Supplemental Table 10 online). For qRT-PCR, the cDNAs were amplified using specific primers and the Full Velocity SYBR-Green QPCR Master Mix (Stratagene) on the Mx3000P thermocycler (Stratagene). Gene-specific primers for select genes were designed with the help of AtRTPrimer, a database for generating specific RT-PCR primer pairs (Han and Kim, 2006), and are shown in Supplemental Table 10 online. PCR was optimized, and reactions were performed in triplicate. The transcript level was standardized based on cDNA amplification of reference genes such as ACTIN2/8 and UBIQUITIN10. Relative gene expression data was generated using the 2−ΔΔCt method (Livak and Schmittgen, 2001) using either the wild type as the reference (for basal expression comparisons) or the 0 time point of each plant line as the reference (for drought-induced expression).

Accession Numbers

Sequence data from the genes studied in this article can be found in the Arabidopsis Genome Initiative or GenBank/EMBL databases under the following accession numbers: RAB18, At5g66400; KIN2, At5g15970; COR15a, At2g42540; ERD1, At5g51070; ACTIN2, At3g18780; UBIQUITIN10, At4g05320; CRT3, At1g08450; BiP3, At1g09080; PR-1, At2g14610; DREB2A, At5g05410; Xero2, At3g50970; GolS2, At1g56600; GolS3, At1g09350; PLC7, At3g55940; Thm, At4g36010; LEA2g, At2g42560; OXG2, At5g43450; ACS2, At1g01480; AGP1, At5g64310; TCH4, At5g57560.

Supplemental Data

The following materials are available in the online version of this article.

  • Supplemental Figure 1. Comparison of Root Growth in Drought-Stressed Plants.

  • Supplemental Figure 2. Detection of InsP 5-ptase Protein in Epidermal Peels.

  • Supplemental Figure 3. Basal ABA Levels in Wild-Type and Transgenic Plants.

  • Supplemental Figure 4. Recovery of Transgenic Plants after Rewatering.

  • Supplemental Figure 5. GO Annotation for Transcripts Commonly Upregulated and Downregulated (Greater Than Twofold) with Drought.

  • Supplemental Table 1. Transcripts Showing Reduced Basal Expression in InsP 5-ptase Plants Compared with the Wild Type and C2 (62 Transcripts).

  • Supplemental Table 2. Transcripts Showing Increased Basal Expression in InsP 5-ptase Plants Compared with the Wild Type and C2 (20 Transcripts).

  • Supplemental Table 3. Transcripts Upregulated (Greater Than Twofold) with Drought in the Wild Type and C2 Only (244 Transcripts).

  • Supplemental Table 4. Transcripts Upregulated with Drought (Greater Than Twofold) in InsP 5-ptase Plants Only (Four Transcripts).

  • Supplemental Table 5. Transcripts Downregulated (Greater Than Twofold) with Drought in the Wild Type and C2 Only (183 Transcripts).

  • Supplemental Table 6. Transcripts Downregulated with Drought (Greater Than Twofold) in InsP 5-ptase Plants Only (22 Transcripts).

  • Supplemental Table 7. Transcripts Commonly Upregulated (Greater Than Twofold) with Drought (219 Transcripts).

  • Supplemental Table 8. Transcripts Commonly Downregulated (Greater Than Twofold) with Drought (161 Transcripts).

  • Supplemental Table 9. Comparison of Basal Levels of Soluble Sugars and Inositol in Wild-Type and Transgenic (T8) Plants.

  • Supplemental Table 10. Primers Used for RT-PCR and qRT-PCR.

Supplementary Material

[Supplemental Data]
[Supplemental Data]

Acknowledgments

We thank Heather and Marc Knight (Durham University) for the Agrobacterium strain containing the aequorin-expressing plasmid and for technical advice on the aequorin measurements. We also thank John York (Duke University) for the use of the HPLC system for the inositol phosphate analysis, Linda Hanley-Bowdoin (Department of Molecular and Structural Biochemistry, North Carolina State University) for the use of the Mx3000P real-time PCR machine, and Bill Hoffmann (Department of Plant Biology, North Carolina State University) for the use of the porometer. We thank Ying-Hsuan Sun (Forest Biotechnology Group, North Carolina State University) for help with sorting the microarray data and former students Peter J. Aspesi, Kelly Althaus, and Diana Pan for technical help. We acknowledge the services of Nigel Deighton and Norm Glassbrook of the Metabolomics and Proteomics Laboratory at North Carolina State University for the ABA measurements and sugar analysis and Expression Analysis for hybridizing and scanning the microarrays and data acquisition. This work was supported by the National Research Initiative of the U.S. Department of Agriculture Cooperative State Research, Education, and Extension Service (Grant 2004-35100-14892 to I.Y.P.) and the National Science Foundation (Grant MCB 0315869 to W.F.B.).

The author responsible for distribution of materials integral to the findings presented in this article in accordance with the policy described in the Instructions for Authors (www.plantcell.org) is: Imara Y. Perera (imara_perera@ncsu.edu).

[W]

Online version contains Web-only data.

References

  1. Agarwal, P., Agarwal, P.K., Nair, S., Sopory, S.K., and Reddy, M.K. (2007). Stress-inducible DREB2A transcription factor from Pennisetum glaucum is a phosphoprotein and its phosphorylation negatively regulates its DNA-binding activity. Mol. Genet. Genomics 277 189–198. [DOI] [PubMed] [Google Scholar]
  2. Bartels, D., and Sunkar, R. (2005). Drought and salt tolerance in plants. Crit. Rev. Plant Sci. 24 23–58. [Google Scholar]
  3. Berdy, S.E., Kudla, J., Gruissem, W., and Gillaspy, G.E. (2001). Molecular characterization of At5PTase1, an inositol phosphatase capable of terminating inositol trisphosphate signaling. Plant Physiol. 126 801–810. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Berridge, M.J. (1993). Inositol trisphosphate and calcium signalling. Nature 361 315–325. [DOI] [PubMed] [Google Scholar]
  5. Boss, W.F., Davis, A.J., Im, Y.J., Galvao, R.M., and Perera, I.Y. (2006). Phosphoinositide metabolism: Towards an understanding of subcellular signaling. Subcell. Biochem. 39 181–205. [DOI] [PubMed] [Google Scholar]
  6. Bray, E.A. (2004). Genes commonly regulated by water-deficit stress in Arabidopsis thaliana. J. Exp. Bot. 55 2331–2341. [DOI] [PubMed] [Google Scholar]
  7. Burnette, R.N., Gunesekera, B.M., and Gillaspy, G.E. (2003). An Arabidopsis inositol 5-phosphatase gain-of-function alters abscisic acid signaling. Plant Physiol. 132 1011–1019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Caird, M.A., Richards, J.H., and Donovan, L.A. (2007). Nighttime stomatal conductance and transpiration in C3 and C4 plants. Plant Physiol. 143 4–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Catala, R., Ouyang, J., Abreu, I.A., Hu, Y., Seo, H., Zhang, X., and Chua, N.H. (2007). The Arabidopsis E3 SUMO ligase SIZ1 regulates plant growth and drought responses. Plant Cell 19 2952–2966. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Christmann, A., Moes, D., Himmelbach, A., Yang, Y., Tang, Y., and Grill, E. (2006). Integration of abscisic acid signalling into plant responses1. Plant Biol. 8 314–325. [DOI] [PubMed] [Google Scholar]
  11. Clarke, J.D., and Zhu, T. (2006). Microarray analysis of the transcriptome as a stepping stone towards understanding biological systems: Practical considerations and perspectives. Plant J. 45 630–650. [DOI] [PubMed] [Google Scholar]
  12. Cockcroft, S., and De Matteis, M.A. (2001). Inositol lipids as spatial regulators of membrane traffic. J. Membr. Biol. 180 187–194. [DOI] [PubMed] [Google Scholar]
  13. De Smedt, F., Missiaen, L., Parys, J.B., Vanweyenberg, V., De Smedt, H., and Erneux, C. (1997). Isoprenylated human brain type I inositol 1,4,5-trisphosphate 5-phosphatase controls Ca2+ oscillations induced by ATP in Chinese hamster ovary cells. J. Biol. Chem. 272 17367–17375. [DOI] [PubMed] [Google Scholar]
  14. DeWald, D.B., Torabinejad, J., Jones, C.A., Shope, J.C., Cangelosi, A.R., Thompson, J.E., Prestwich, G.D., and Hama, H. (2001). Rapid accumulation of phosphatidylinositol 4,5-bisphosphate and inositol 1,4,5-trisphosphate correlates with calcium mobilization in salt-stressed Arabidopsis. Plant Physiol. 126 759–769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Di Paolo, G., and De Camilli, P. (2006). Phosphoinositides in cell regulation and membrane dynamics. Nature 443 651–657. [DOI] [PubMed] [Google Scholar]
  16. Engstrom, E.M., Ehrhardt, D.W., Mitra, R.M., and Long, S.R. (2002). Pharmacological analysis of nod factor-induced calcium spiking in Medicago truncatula. Evidence for the requirement of type IIA calcium pumps and phosphoinositide signaling. Plant Physiol. 128 1390–1401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Ercetin, M.E., and Gillaspy, G.E. (2004). Molecular characterization of an Arabidopsis gene encoding a phospholipid-specific inositol polyphosphate 5-phosphatase. Plant Physiol. 135 938–946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Franklin-Tong, V.E., Drobak, B.K., Allan, A.C., Watkins, P., and Trewavas, A.J. (1996). Growth of pollen tubes of Papaver rhoeas is regulated by a slow-moving calcium wave propagated by inositol 1,4,5-trisphosphate. Plant Cell 8 1305–1321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Gilroy, S., Read, N.D., and Trewavas, A.J. (1990). Elevation of cytoplasmic calcium by caged calcium or caged inositol triphosphate initiates stomatal closure. Nature 346 769–771. [DOI] [PubMed] [Google Scholar]
  20. Gong, Q., Li, P., Ma, S., Indu Rupassara, S., and Bohnert, H.J. (2005). Salinity stress adaptation competence in the extremophile Thellungiella halophila in comparison with its relative Arabidopsis thaliana. Plant J. 44 826–839. [DOI] [PubMed] [Google Scholar]
  21. Griffiths, A., and Bray, E.A. (1996). Shoot induction of ABA-requiring genes in response to soil drying. J. Exp. Bot. 47 1525–1531. [Google Scholar]
  22. Gunesekera, B., Torabinejad, J., Robinson, J., and Gillaspy, G.E. (2007). Inositol polyphosphate 5-phosphatases 1 and 2 are required for regulating seedling growth. Plant Physiol. 143 1408–1417. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Han, S., and Kim, D. (2006). AtRTPrimer: Database for Arabidopsis genome-wide homogeneous and specific RT-PCR primer-pairs. BMC Bioinformatics 7 179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Hetherington, A.M., and Brownlee, C. (2004). The generation of Ca2+ signals in plants. Annu. Rev. Plant Biol. 55 401–427. [DOI] [PubMed] [Google Scholar]
  25. Hitz, W.D., Carlson, T.J., Kerr, P.S., and Sebastian, S.A. (2002). Biochemical and molecular characterization of a mutation that confers a decreased raffinosaccharide and phytic acid phenotype on soybean seeds. Plant Physiol. 128 650–660. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Hundertmark, M., and Hincha, D. (2008). LEA (Late Embryogenesis Abundant) proteins and their encoding genes in Arabidopsis thaliana. BMC Genomics 9 118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Hunt, L., Mills, L.N., Pical, C., Leckie, C.P., Aitken, F.L., Kopka, J., Mueller-Roeber, B., McAinsh, M.R., Hetherington, A.M., and Gray, J.E. (2003). Phospholipase C is required for the control of stomatal aperture by ABA. Plant J. 34 47–55. [DOI] [PubMed] [Google Scholar]
  28. Hunt, L., Otterhag, L., Lee, J.C., Lasheen, T., Hunt, J., Seki, M., Shinozaki, K., Sommarin, M., Gilmour, D.J., Pical, C., and Gray, J.E. (2004). Gene-specific expression and calcium activation of Arabidopsis thaliana phospholipase C isoforms. New Phytol. 162 643–654. [DOI] [PubMed] [Google Scholar]
  29. Im, Y.J., Davis, A.J., Perera, I.Y., Johannes, E., Allen, N.S., and Boss, W.F. (2007). The N-terminal membrane occupation and recognition nexus domain of Arabidopsis phosphatidylinositol phosphate kinase 1 regulates enzyme activity. J. Biol. Chem. 282 5443–5452. [DOI] [PubMed] [Google Scholar]
  30. Israelsson, M., Siegel, R.S., Young, J., Hashimoto, M., Iba, K., and Schroeder, J.I. (2006). Guard cell ABA and CO2 signaling network updates and Ca2+ sensor priming hypothesis. Curr. Opin. Plant Biol. 9 654–663. [DOI] [PubMed] [Google Scholar]
  31. Jung, J.-Y., Kim, Y.-W., Kwak, J.M., Hwang, J.-U., Young, J., Schroeder, J.I., Hwang, I., and Lee, Y. (2002). Phosphatidylinositol 3- and 4-phosphate are required for normal stomatal movements. Plant Cell 14 2399–2412. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Karner, U., Peterbauer, T., Raboy, V., Jones, D.A., Hedley, C.L., and Richter, A. (2004). myo-Inositol and sucrose concentrations affect the accumulation of raffinose family oligosaccharides in seeds. J. Exp. Bot. 55 1981–1987. [DOI] [PubMed] [Google Scholar]
  33. Kasuga, M., Liu, Q., Miura, S., Yamaguchi-Shinozaki, K., and Shinozaki, K. (1999). Improving plant drought, salt, and freezing tolerance by gene transfer of a single stress-inducible transcription factor. Nat. Biotechnol. 17 287–291. [DOI] [PubMed] [Google Scholar]
  34. Kawaguchi, R., Girke, T., Bray, E.A., and Bailey-Serres, J. (2004). Differential mRNA translation contributes to gene regulation under non-stress and dehydration stress conditions in Arabidopsis thaliana. Plant J. 38 823–839. [DOI] [PubMed] [Google Scholar]
  35. Kawaguchi, R., Williams, A.J., Bray, E.A., and Bailey-Serres, J. (2003). Water-deficit-induced translational control in Nicotiana tabacum. Plant Cell Environ. 26 221–229. [Google Scholar]
  36. Knight, H., and Knight, M.R. (2001). Abiotic stress signalling pathways: Specificity and cross-talk. Trends Plant Sci. 6 262–267. [DOI] [PubMed] [Google Scholar]
  37. Knight, H., Trewavas, A.J., and Knight, M.R. (1996). Cold calcium signaling in Arabidopsis involves two cellular pools and a change in calcium signature after acclimation. Plant Cell 8 489–503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Knight, H., Trewavas, A.J., and Knight, M.R. (1997. a). Calcium signalling in Arabidopsis thaliana responding to drought and salinity. Plant J. 12 1067–1078. [DOI] [PubMed] [Google Scholar]
  39. Knight, H., Trewavas, A.J., and Knight, M.R. (1997. b). Recombinant aequorin methods for measurement of intracellular calcium in plants. Plant Mol. Biol. Manual C4 1–22. [Google Scholar]
  40. Kreps, J.A., Wu, Y., Chang, H.-S., Zhu, T., Wang, X., and Harper, J.F. (2002). Transcriptome changes for Arabidopsis in response to salt, osmotic, and cold stress. Plant Physiol. 130 2129–2141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Krinke, O., Novotna, Z., Valentova, O., and Martinec, J. (2007). Inositol trisphosphate receptor in higher plants: Is it real? J. Exp. Bot. 58 361–376. [DOI] [PubMed] [Google Scholar]
  42. Kubitscheck, U., Homann, U., and Thiel, G. (2000). Osmotically evoked shrinking of guard-cell protoplasts causes vesicular retrieval of plasma membrane into the cytoplasm. Planta 210 423–431. [DOI] [PubMed] [Google Scholar]
  43. Laxminarayan, K.M., Chan, B.K., Tetaz, T., Bird, P.I., and Mitchell, C.A. (1994). Characterization of a cDNA encoding the 43-kDa membrane-associated inositol-polyphosphate 5-phosphatase. J. Biol. Chem. 269 17305–17310. [PubMed] [Google Scholar]
  44. Laxminarayan, K.M., Matzaris, M., Speed, C.J., and Mitchell, C.A. (1993). Purification and characterization of a 43-kDa membrane-associated inositol polyphosphate 5-phosphatase from human placenta. J. Biol. Chem. 268 4968–4974. [PubMed] [Google Scholar]
  45. Lee, Y., Choi, Y.B., Suh, S., Lee, J., Assmann, S.M., Joe, C.O., Kelleher, J.F., and Crain, R.C. (1996). Abscisic acid-induced phosphoinositide turnover in guard cell protoplasts of Vicia faba. Plant Physiol. 110 987–996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Lee, Y., Kim, Y.W., Jeon, B.W., Park, K.Y., Suh, S.J., Seo, J., Kwak, J.M., Martinoia, E., Hwang, I., and Lee, Y. (2007). Phosphatidylinositol 4,5-bisphosphate is important for stomatal opening. Plant J. 52 803–816. [DOI] [PubMed] [Google Scholar]
  47. Lemtiri-Chlieh, F., MacRobbie, E.A., and Brearley, C.A. (2000). Inositol hexakisphosphate is a physiological signal regulating the K+-inward rectifying conductance in guard cells. Proc. Natl. Acad. Sci. USA 97 8687–8692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Lemtiri-Chlieh, F., MacRobbie, E.A., Webb, A.A., Manison, N.F., Brownlee, C., Skepper, J.N., Chen, J., Prestwich, G.D., and Brearley, C.A. (2003). Inositol hexakisphosphate mobilizes an endomembrane store of calcium in guard cells. Proc. Natl. Acad. Sci. USA 100 10091–10095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Levchenko, V., Konrad, K.R., Dietrich, P., Roelfsema, M.R.G., and Hedrich, R. (2005). Cytosolic abscisic acid activates guard cell anion channels without preceding Ca2+ signals. Proc. Natl. Acad. Sci. USA 102 4203–4208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Li, S., Assmann, S.M., and Albert, R. (2006. a). Predicting essential components of signal transduction networks: A dynamic model of guard cell abscisic acid signaling. PLoS Biol. 4 e312. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Li, Y., Lee, K.K., Walsh, S., Smith, C., Hadingham, S., Sorefan, K., Cawley, G., and Bevan, M.W. (2006. b). Establishing glucose- and ABA-regulated transcription networks in Arabidopsis by microarray analysis and promoter classification using a relevance vector machine. Genome Res. 16 414–427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Liu, Q., Kasuga, M., Sakuma, Y., Abe, H., Miura, S., Yamaguchi-Shinozaki, K., and Shinozaki, K. (1998). Two transcription factors, DREB1 and DREB2, with an EREBP/AP2 DNA binding domain separate two cellular signal transduction pathways in drought- and low-temperature-responsive gene expression, respectively, in Arabidopsis. Plant Cell 10 1391–1406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Livak, K.J., and Schmittgen, T.D. (2001). Analysis of relative gene expression data using real-time quantitative PCR and the 2−ΔΔCT method. Methods 25 402–408. [DOI] [PubMed] [Google Scholar]
  54. Majerus, P.W., Kisseleva, M.V., and Norris, F.A. (1999). The role of phosphatases in inositol signaling reactions. J. Biol. Chem. 274 10669–10672. [DOI] [PubMed] [Google Scholar]
  55. Martin, T.F.J. (1998). Phosphoinositide lipids as signaling molecules: Common themes for signal transduction, cytoskeletal regulation, and membrane trafficking. Annu. Rev. Cell Dev. Biol. 14 231–264. [DOI] [PubMed] [Google Scholar]
  56. Meijer, H.J., and Munnik, T. (2003). Phospholipid-based signaling in plants. Annu. Rev. Plant Biol. 54 265–306. [DOI] [PubMed] [Google Scholar]
  57. Mills, L.N., Hunt, L., Leckie, C.P., Aitken, F.L., Wentworth, M., McAinsh, M.R., Gray, J.E., and Hetherington, A.M. (2004). The effects of manipulating phospholipase C on guard cell ABA-signalling. J. Exp. Bot. 55 199–204. [DOI] [PubMed] [Google Scholar]
  58. Mishra, G., Zhang, W., Deng, F., Zhao, J., and Wang, X. (2006). A bifurcating pathway directs abscisic acid effects on stomatal closure and opening in Arabidopsis. Science 312 264–266. [DOI] [PubMed] [Google Scholar]
  59. Mitchell, C.A., Brown, S., Campbell, J.K., Munday, A.D., and Speed, C.J. (1996). Regulation of second messengers by the inositol polyphosphate 5-phosphatases. Biochem. Soc. Trans. 24 994–1000. [DOI] [PubMed] [Google Scholar]
  60. Mueller-Roeber, B., and Pical, C. (2002). Inositol phospholipid metabolism in Arabidopsis. Characterized and putative isoforms of inositol phospholipid kinase and phosphoinositide-specific phospholipase C. Plant Physiol. 130 22–46. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Munnik, T., Meijer, H.J., Ter Riet, B., Hirt, H., Frank, W., Bartels, D., and Musgrave, A. (2000). Hyperosmotic stress stimulates phospholipase D activity and elevates the levels of phosphatidic acid and diacylglycerol pyrophosphate. Plant J. 22 147–154. [DOI] [PubMed] [Google Scholar]
  62. Ng, C.K.Y., McAinsh, M.R., Gray, J.E., Hunt, L., Leckie, C.P., Mills, L., and Hetherington, A.M. (2001). Calcium-based signalling systems in guard cells. New Phytol. 151 109–120. [DOI] [PubMed] [Google Scholar]
  63. Nishizawa, A., Yabuta, Y., Yoshida, E., Maruta, T., Yoshimura, K., and Shigeoka, S. (2006). Arabidopsis heat shock transcription factor A2 as a key regulator in response to several types of environmental stress. Plant J. 48 535–547. [DOI] [PubMed] [Google Scholar]
  64. Obendorf, R.L. (1997). Oligosaccharides and galactosyl cyclitols in seed desiccation tolerance. Seed Sci. Res. 7 63–74. [Google Scholar]
  65. Pawitan, Y., Michiels, S., Koscielny, S., Gusnanto, A., and Ploner, A. (2005). False discovery rate, sensitivity and sample size for microarray studies. Bioinformatics 21 3017–3024. [DOI] [PubMed] [Google Scholar]
  66. Perera, I.Y., Davis, A.J., Galanopoulou, D., Im, Y.J., and Boss, W.F. (2005). Characterization and comparative analysis of Arabidopsis phosphatidylinositol phosphate 5-kinase 10 reveals differences in Arabidopsis and human phosphatidylinositol phosphate kinases. FEBS Lett. 579 3427–3432. [DOI] [PubMed] [Google Scholar]
  67. Perera, I.Y., Heilmann, I., Chang, S.C., Boss, W.F., and Kaufman, P.B. (2001). A role for inositol 1,4,5-trisphosphate in gravitropic signaling and the retention of cold-perceived gravistimulation of oat shoot pulvini. Plant Physiol. 125 1499–1507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Perera, I.Y., Hung, C.Y., Brady, S., Muday, G.K., and Boss, W.F. (2006). A universal role for inositol 1,4,5-trisphosphate-mediated signaling in plant gravitropism. Plant Physiol. 140 746–760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Perera, I.Y., Love, J., Heilmann, I., Thompson, W.F., and Boss, W.F. (2002). Up-regulation of phosphoinositide metabolism in tobacco cells constitutively expressing the human type I inositol polyphosphate 5-phosphatase. Plant Physiol. 129 1795–1806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Qin, C., and Wang, X. (2002). The Arabidopsis phospholipase D family. Characterization of a calcium-independent and phosphatidylcholine-selective PLD zeta 1 with distinct regulatory domains. Plant Physiol. 128 1057–1068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Raboy, V. (2003). myo-Inositol-1,2,3,4,5,6-hexakisphosphate. Phytochemistry 64 1033–1043. [DOI] [PubMed] [Google Scholar]
  72. Roelfsema, M.R., and Hedrich, R. (2005). In the light of stomatal opening: New insights into ‘the Watergate.’ New Phytol. 167 665–691. [DOI] [PubMed] [Google Scholar]
  73. Roelfsema, M.R.G., and Prins, H.B.A. (1995). Effect of abscisic acid on stomatal opening in isolated epidermal strips of abi mutants of Arabidopsis thaliana. Physiol. Plant. 95 373–378. [Google Scholar]
  74. Sakuma, Y., Maruyama, K., Osakabe, Y., Qin, F., Seki, M., Shinozaki, K., and Yamaguchi-Shinozaki, K. (2006. a). Functional analysis of an Arabidopsis transcription factor, DREB2A, involved in drought-responsive gene expression. Plant Cell 18 1292–1309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Sakuma, Y., Maruyama, K., Qin, F., Osakabe, Y., Shinozaki, K., and Yamaguchi-Shinozaki, K. (2006. b). Colloquium Paper. Dual function of an Arabidopsis transcription factor DREB2A in water-stress-responsive and heat-stress-responsive gene expression. Proc. Natl. Acad. Sci. USA 103 18822–18827. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Sanchez, J.-P., and Chua, N.-H. (2001). Arabidopsis PLC1 is required for secondary responses to abscisic acid signals. Plant Cell 13 1143–1154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Sanders, D., Brownlee, C., and Harper, J.F. (1999). Communicating with calcium. Plant Cell 11 691–706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Sanders, D., Pelloux, J., Brownlee, C., and Harper, J.F. (2002). Calcium at the crossroads of signaling. Plant Cell 14 (suppl.): S401–S417. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Schramm, F., Larkindale, J., Kiehlmann, E., Ganguli, A., Englich, G., Vierling, E., and von Koskull-Doring, P. (2008). A cascade of transcription factor DREB2A and heat stress transcription factor HsfA3 regulates the heat stress response of Arabidopsis. Plant J. 53 264–274. [DOI] [PubMed] [Google Scholar]
  80. Schroeder, J.I., Allen, G.J., Hugouvieux, V., Kwak, J.M., and Waner, D. (2001). Guard cell signal transduction. Annu. Rev. Plant Physiol. Plant Mol. Biol. 52 627–658. [DOI] [PubMed] [Google Scholar]
  81. Seki, M., Narusaka, M., Abe, H., Kasuga, M., Yamaguchi-Shinozaki, K., Carninci, P., Hayashizaki, Y., and Shinozaki, K. (2001). Monitoring the expression pattern of 1300 Arabidopsis genes under drought and cold stresses by using a full-length cDNA microarray. Plant Cell 13 61–72. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Shinozaki, K., and Yamaguchi-Shinozaki, K. (2000). Molecular responses to dehydration and low temperature: Differences and cross-talk between two stress signaling pathways. Curr. Opin. Plant Biol. 3 217–223. [PubMed] [Google Scholar]
  83. Shinozaki, K., and Yamaguchi-Shinozaki, K. (2007). Gene networks involved in drought stress response and tolerance. J. Exp. Bot. 58 221–227. [DOI] [PubMed] [Google Scholar]
  84. Staxen, I.I., Pical, C., Montgomery, L.T., Gray, J.E., Hetherington, A.M., and McAinsh, M.R. (1999). Abscisic acid induces oscillations in guard-cell cytosolic free calcium that involve phosphoinositide-specific phospholipase C. Proc. Natl. Acad. Sci. USA 96 1779–1784. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Stevenson, J.M., Perera, I.Y., Heilmann, I.I., Persson, S., and Boss, W.F. (2000). Inositol signaling and plant growth. Trends Plant Sci. 5 252–258. [DOI] [PubMed] [Google Scholar]
  86. Stevenson-Paulik, J., Bastidas, R.J., Chiou, S.-T., Frye, R.A., and York, J.D. (2005). Generation of phytate-free seeds in Arabidopsis through disruption of inositol polyphosphate kinases. Proc. Natl. Acad. Sci. USA 102 12612–12617. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Stevenson-Paulik, J., Odom, A.R., and York, J.D. (2002). Molecular and biochemical characterization of two plant inositol polyphosphate 6-/3-/5-kinases. J. Biol. Chem. 277 42711–42718. [DOI] [PubMed] [Google Scholar]
  88. Taji, T., Ohsumi, C., Iuchi, S., Seki, M., Kasuga, M., Kobayashi, M., Yamaguchi-Shinozaki, K., and Shinozaki, K. (2002). Important roles of drought- and cold-inducible genes for galactinol synthase in stress tolerance in Arabidopsis thaliana. Plant J. 29 417–426. [DOI] [PubMed] [Google Scholar]
  89. Taji, T., Seki, M., Satou, M., Sakurai, T., Kobayashi, M., Ishiyama, K., Narusaka, Y., Narusaka, M., Zhu, J.K., and Shinozaki, K. (2004). Comparative genomics in salt tolerance between Arabidopsis and Arabidopsis-related halophyte salt cress using Arabidopsis microarray. Plant Physiol. 135 1697–1709. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Tang, R.H., Han, S., Zheng, H., Cook, C.W., Choi, C.S., Woerner, T.E., Jackson, R.B., and Pei, Z.M. (2007). Coupling diurnal cytosolic Ca2+ oscillations to the CAS-IP3 pathway in Arabidopsis. Science 315 1423–1426. [DOI] [PubMed] [Google Scholar]
  91. Testerink, C., and Munnik, T. (2005). Phosphatidic acid: A multifunctional stress signaling lipid in plants. Trends Plant Sci. 10 368–375. [DOI] [PubMed] [Google Scholar]
  92. Thomashow, M. (1999). Plant cold acclimation: Freezing tolerance genes and regulatory mechanism. Annu. Rev. Plant Physiol. Plant Mol. Biol. 50 571–599. [DOI] [PubMed] [Google Scholar]
  93. Tucker, E.B., and Boss, W.F. (1996). Mastoparan-induced intracellular Ca2+ fluxes may regulate cell-to-cell communication in plants. Plant Physiol. 111 459–467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Valliyodan, B., and Nguyen, H.T. (2006). Understanding regulatory networks and engineering for enhanced drought tolerance in plants. Curr. Opin. Plant Biol. 9 189–195. [DOI] [PubMed] [Google Scholar]
  95. Wang, X. (2004). Lipid signaling. Curr. Opin. Plant Biol. 7 329–336. [DOI] [PubMed] [Google Scholar]
  96. Weigel, D., and Glazebrook, J. (2002). Arabidopsis: A Laboratory Manual. (Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press).
  97. Williams, M.E., Torabinejad, J., Cohick, E., Parker, K., Drake, E.J., Thompson, J.E., Hortter, M., and Dewald, D.B. (2005). Mutations in the Arabidopsis phosphoinositide phosphatase gene SAC9 lead to overaccumulation of PtdIns(4,5)P2 and constitutive expression of the stress-response pathway. Plant Physiol. 138 686–700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Xia, H.J., Brearley, C., Elge, S., Kaplan, B., Fromm, H., and Mueller-Roeber, B. (2003). Arabidopsis inositol polyphosphate 6-/3-kinase is a nuclear protein that complements a yeast mutant lacking a functional ArgR-Mcm1 transcription complex. Plant Cell 15 449–463. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Yamaguchi-Shinozaki, K., and Shinozaki, K. (2006). Transcriptional regulatory networks in cellular responses and tolerance to dehydration and cold stresses. Annu. Rev. Plant Biol. 57 781–803. [DOI] [PubMed] [Google Scholar]
  100. Zhang, W., Qin, C., Zhao, J., and Wang, X. (2004). Phospholipase D alpha 1-derived phosphatidic acid interacts with ABI1 phosphatase 2C and regulates abscisic acid signaling. Proc. Natl. Acad. Sci. USA 101 9508–9513. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Zhang, W., Yu, L., Zhang, Y., and Wang, X. (2005). Phospholipase D in the signaling networks of plant response to abscisic acid and reactive oxygen species. Biochim. Biophys. Acta 1736 1–9. [DOI] [PubMed] [Google Scholar]
  102. Zhu, J.K. (2002). Salt and drought stress signal transduction in plants. Annu. Rev. Plant Biol. 53 247–273. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

[Supplemental Data]
[Supplemental Data]

Articles from The Plant Cell are provided here courtesy of Oxford University Press

RESOURCES