首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Stomata control the exchange of CO2 and water vapor in land plants. Thus, whereas a constant supply of CO2 is required to maintain adequate rates of photosynthesis, the accompanying water losses must be tightly regulated to prevent dehydration and undesired metabolic changes. Accordingly, the uptake or release of ions and metabolites from guard cells is necessary to achieve normal stomatal function. The AtQUAC1, an R-type anion channel responsible for the release of malate from guard cells, is essential for efficient stomatal closure. Here, we demonstrate that mutant plants lacking AtQUAC1 accumulated higher levels of malate and fumarate. These mutant plants not only display slower stomatal closure in response to increased CO2 concentration and dark but are also characterized by improved mesophyll conductance. These responses were accompanied by increases in both photosynthesis and respiration rates, without affecting the activity of photosynthetic and respiratory enzymes and the expression of other transporter genes in guard cells, which ultimately led to improved growth. Collectively, our results highlight that the transport of organic acids plays a key role in plant cell metabolism and demonstrate that AtQUAC1 reduce diffusive limitations to photosynthesis, which, at least partially, explain the observed increments in growth under well-watered conditions.Stomata are functionally specialized microscopic pores that control the essential exchange of CO2 and H2O with the environment in land plants. Stomata are found on the surfaces of the majority of the aerial parts of plants, rendering them as the main control point regulating the flow of gases between plants and their surrounding atmosphere. Accordingly, the majority of water loss from plants occurs through stomatal pores, allowing plant transpiration and CO2 absorption for the photosynthetic process (Bergmann and Sack, 2007; Kim et al., 2010). The maintenance of an adequate water balance through stomatal control is crucial to plants because cell expansion and growth require tissues to remain turgid (Sablowski and Carnier Dornelas, 2014), and minor reductions in cell water volume and turgor pressure will therefore compromise both processes (Thompson, 2005). As a result, the high sensitivity of plant tissues to turgor has prompted the use of reverse genetic studies in attempt to engineer plants with improved performance (Cowan and Troughton, 1971; Xiong et al., 2009; Borland et al., 2014; Franks et al., 2015).In most land plants, not only redox signals invoked by shifts in light quality (Busch, 2014) but also the transport of inorganic ions (e.g. K+, Cl, and NO3) as well as metabolites such as the phytohormone abscisic acid (ABA), Suc, and malate, are important players controlling stomatal movements (Hetherington, 2001; Roelfsema and Hedrich, 2005; Pandey et al., 2007; Blatt et al., 2014; Kollist et al., 2014). In this context, although organic acids in plants is known to support numerous and diverse functions both within and beyond cellular metabolism, only recently have we obtained genetic evidence to support that modulation of guard cell malate and fumarate concentration can greatly influence stomatal movements (Nunes-Nesi et al., 2007; Araújo et al., 2011b; Penfield et al., 2012; Medeiros et al., 2015). Notably malate, in particular, has been considered as a key metabolite and one of the most important organic metabolites involved in guard cell movements (Hedrich and Marten, 1993; Fernie and Martinoia, 2009; Meyer et al., 2010). During stomatal aperture, the flux of malate into guard cells coupled with hexoses generated on starch breakdown lead to decreases in the water potential, and consequently, water uptake by the guard cells ultimately opens the stomata pore (Roelfsema and Hedrich, 2005; Vavasseur and Raghavendra, 2005; Lee et al., 2008). On the other hand, during stomatal closure, malate is believed to be converted into starch, which has no osmotic activity (Penfield et al., 2012) or, alternatively, is released from the guard cells to the surrounding apoplastic space (Lee et al., 2008; Negi et al., 2008; Vahisalu et al., 2008; Meyer et al., 2010).The role of organic acids on the stomatal movements has been largely demonstrated by studies related to malate transport (Lee et al., 2008; Meyer et al., 2010; Sasaki et al., 2010). In the last decade, two protein families were identified and functionally characterized to be directly involved with organic acid transport at the guard cell plasma membrane and to be required for stomatal functioning (Lee et al., 2008; Meyer et al., 2010; Sasaki et al., 2010). In summary, AtABCB14, a member of the ABC (ATP binding cassette) family, which is involved in malate transport from apoplast to guard cells, was described as a negative modulator of stomatal closure induced by high CO2 concentration; notably, exogenous application of malate minimizes this response (Lee et al., 2008). In addition, members of a small gene family, which encode the anion channels SLAC1 (slow anion channel 1) and four SLAC1-homologs (SLAHs) in Arabidopsis (Arabidopsis thaliana), have been described to be involved in stomatal movements. SLAC1 is a well-documented S-type anion channel that preferentially transports chloride and nitrate as opposed to malate (Vahisalu et al., 2008, 2010; Geiger et al., 2010; Du et al., 2011; Brandt et al., 2012; Kusumi et al., 2012). Lack of SLAC1 in Arabidopsis and rice (Oryza sativa) culminated in a failure in stomatal closure in response to high CO2 levels, low relative humidity, and dark conditions (Negi et al., 2008; Vahisalu et al., 2008; Kusumi et al., 2012). Although mutations in AtSLAC1 impair S-type anion channel functions as a whole, the R-type anion channel remained functional (Vahisalu et al., 2008). Indeed, a member of the aluminum-activated malate transporter (ALMT) family, AtALMT12, an R-type anion channel, has been demonstrated to be involved in malate transport, particularly at the plasma membrane of guard cells (Meyer et al., 2010; Sasaki et al., 2010). Although AtALMT12 is a member of ALMT family, it is not activated by aluminum, and therefore Meyer et al. (2010) proposed to rename it as AtQUAC1 (quick-activating anion channel 1; Imes et al., 2013; Mumm et al., 2013). Hereafter, we will follow this nomenclature. Deficiency of a functional AtQUAC1 has been documented to lead to changes in stomatal closure in response to high levels of CO2, dark, and ABA (Meyer et al., 2010). Taken together, these studies have clearly demonstrated that both S- and R-type anion channels are key modulators of stomatal movements in response to several environmental factors.Despite a vast number of studies involving the above-mentioned anion channels, little information concerning the metabolic changes caused by their impairment is currently available. Such information is important to understand stomatal movements, mainly considering that organic acids, especially the levels of malate in apoplastic/mesophyll cells, have been highlighted as of key importance in leaf metabolism (Fernie and Martinoia, 2009; Araújo et al., 2011a, 2011b; Lawson et al., 2014; Medeiros et al., 2015). Here, we demonstrate that a disruption in the expression of AtQUAC1, which leads to impaired stomatal closure (Meyer et al., 2010), was accompanied by increases in mesophyll conductance (gm), which is defined as the conductance for the transfer of CO2 from the intercellular airspaces (Ci) to the sites of carboxylation in the chloroplastic stroma (Cc). By further characterization of atquac1 knockout plants, we demonstrated that reduced diffusive limitations resulted in higher photosynthetic rates and altered respiration that, in turn, led to enhanced biomass accumulation. Overall, the results obtained are discussed both in terms of the importance of organic acid transport in plant cell metabolism and with regard to the contribution that it plays in the regulation of both stomatal function and growth.  相似文献   

2.
Light is a major environmental factor required for stomatal opening. Blue light (BL) induces stomatal opening in higher plants as a signal under the photosynthetic active radiation. The stomatal BL response is not present in the fern species of Polypodiopsida. The acquisition of a stomatal BL response might provide competitive advantages in both the uptake of CO2 and prevention of water loss with the ability to rapidly open and close stomata. We surveyed the stomatal opening in response to strong red light (RL) and weak BL under the RL with gas exchange technique in a diverse selection of plant species from euphyllophytes, including spermatophytes and monilophytes, to lycophytes. We showed the presence of RL-induced stomatal opening in most of these species and found that the BL responses operated in all euphyllophytes except Polypodiopsida. We also confirmed that the stomatal opening in lycophytes, the early vascular plants, is driven by plasma membrane proton-translocating adenosine triphosphatase and K+ accumulation in guard cells, which is the same mechanism operating in stomata of angiosperms. These results suggest that the early vascular plants respond to both RL and BL and actively regulate stomatal aperture. We also found three plant species that absolutely require BL for both stomatal opening and photosynthetic CO2 fixation, including a gymnosperm, C. revoluta, and the ferns Equisetum hyemale and Psilotum nudum.Stomata regulate gas exchange between plants and the atmosphere (Zeiger, 1983; Assmann, 1993; Roelfsema and Hedrich, 2005; Shimazaki et al., 2007; Kim et al., 2010). Acquisition of stomata was a key step in the evolution of terrestrial plants by allowing uptake of CO2 from the atmosphere and accelerating the provision of nutrients via the transpiration stream within the plant (Hetherington and Woodward, 2003; McAdam and Brodribb, 2013). Stomatal aperture is regulated by changes in the turgor of guard cells, which are induced by environmental factors and endogenous phytohormones. Light is a major factor in the promotion of stomatal opening, and the opening is mediated via two distinct light-regulated pathways that are known as photosynthesis- and blue light (BL)-dependent responses under photosynthetic active radiation (PAR; Vavasseur and Raghavendra, 2005; Shimazaki et al., 2007; Lawson et al., 2014).The photosynthesis-dependent stomatal opening is induced by a continuous high intensity of light, and the action spectrum for the stomatal opening resembles that of photosynthetic pigments in leaves (Willmer and Fricker, 1996). Both mesophyll and guard cells possess photosynthetically active chloroplasts, and their photosynthesis has been suggested to contribute to stomatal opening in leaves. The decrease in the concentration of intercellular CO2 (Ci) caused by photosynthetic CO2 fixation or some unidentified mediators and metabolites from mesophyll cells is supposed to elicit stomatal opening, although the exact nature of the events is unclear (Wong et al., 1979; Vavasseur and Raghavendra, 2005; Roelfsema et al., 2006; Mott et al., 2008; Lawson et al., 2014).BL-dependent stomatal opening requires a strong intensity of PAR as a background: weak BL solely scarcely elicits stomatal opening, but the same intensity of BL induces the fast and large stomatal opening in the presence of strong red light (RL; Ogawa et al., 1978; Shimazaki et al., 2007). Since such stomatal opening requires BL under the RL or PAR, we call the opening reaction a BL-dependent response of stomata. BL-dependent stomatal response takes place and proceeds in natural environments because the sunlight contains both BL and RL and facilitates photosynthetic CO2 fixation (Assmann, 1988; Takemiya et al., 2013a). In this stomatal response, BL and PAR (BL, RL, and other wavelengths of light) seem to act as a signal and an energy source, respectively.The BL-dependent stomatal opening is initiated by the absorption of BL by phototropin1 and phototropin2 (Kinoshita et al., 2001), the plant-specific BL receptors, in guard cells followed by activation of the plasma membrane proton-translocating adenosine triphosphatase (H+-ATPase; Kinoshita and Shimazaki, 1999). Two newly identified proteins, protein phosphatase1 and blue light signaling1 (BLUS1), mediate the signaling between phototropins and H+-ATPase (Takemiya et al., 2006, 2013a, 2013b). The activated H+-ATPase evokes a plasma membrane hyperpolarization, which drives K+ uptake through the voltage-gated, inward-rectifying K+ channels (Assmann, 1993; Shimazaki et al., 2007; Kim et al., 2010; Kollist et al., 2014). The accumulation of K+ causes water uptake and increases turgor pressure of guard cells, and finally results in stomatal opening. The BL-dependent opening is enhanced by RL, and BL at low intensity is effective in the presence of RL (Ogawa et al., 1978; Iino et al., 1985; Shimazaki et al., 2007; Suetsugu et al., 2014). These stomatal responses by RL and BL are commonly observed in a number of seed plants so far examined.Fine control of stomatal aperture to various environmental factors has been characterized in many angiosperms. Although morphological and mechanical diversity of stomata is widely documented, little is known about the functional diversity (Willmer and Fricker, 1996; Hetherington and Woodward, 2003). Our previous study indicated that BL-dependent stomatal response is absent in the major fern species of Polypodiopsida, including Adiantum capillus-veneris, Pteris cretica, Asplenium scolopendrium, and Nephrolepis auriculata, but the stomata of these species open by PAR including RL (Doi et al., 2006). When the epidermal peels isolated from A. capillus-veneris are treated with photosynthetic electron transport inhibitor 3-(3,4-dichlorophenyl)-1,1dimethylurea (Doi and Shimazaki, 2008), the response is completely inhibited, but the responses in the seed plants of Vicia faba and Commelina communis are relatively insensitive to 3-(3,4-dichlorophenyl)-1,1dimethylurea (Schwartz and Zeiger, 1984). These findings suggest that there is functional diversity in light-dependent stomatal response in different lineages of land plants. In accord with this notion, the different sensitivities of stomatal response to abscisic acid and CO2 have been reported among the plant species of angiosperm, gymnosperm, ferns, and lycophytes (Mansfield and Willmer, 1969; Brodribb and McAdam, 2011), although the exact responsiveness to abscisic acid and CO2 is still debated (Chater et al., 2011, 2013; Ruszala et al., 2011; McAdam and Brodribb, 2013).To address the origin and distribution of stomatal light responses, we investigated the presence of a stomatal response using a gas exchange method and various lineages of vascular plants, including euphyllophytes and lycophytes. Unexpectedly, all plant lineages except Polypodiopsida in monilophytes exhibited a stomatal response to BL in the presence of RL, suggesting that the response was present in the early evolutionary stage of vascular plants. We also report the stomatal opening in response to RL in these plant species.  相似文献   

3.
Cytosolic Ca2+ in guard cells plays an important role in stomatal movement responses to environmental stimuli. These cytosolic Ca2+ increases result from Ca2+ influx through Ca2+-permeable channels in the plasma membrane and Ca2+ release from intracellular organelles in guard cells. However, the genes encoding defined plasma membrane Ca2+-permeable channel activity remain unknown in guard cells and, with some exceptions, largely unknown in higher plant cells. Here, we report the identification of two Arabidopsis (Arabidopsis thaliana) cation channel genes, CNGC5 and CNGC6, that are highly expressed in guard cells. Cytosolic application of cyclic GMP (cGMP) and extracellularly applied membrane-permeable 8-Bromoguanosine 3′,5′-cyclic monophosphate-cGMP both activated hyperpolarization-induced inward-conducting currents in wild-type guard cells using Mg2+ as the main charge carrier. The cGMP-activated currents were strongly blocked by lanthanum and gadolinium and also conducted Ba2+, Ca2+, and Na+ ions. cngc5 cngc6 double mutant guard cells exhibited dramatically impaired cGMP-activated currents. In contrast, mutations in CNGC1, CNGC2, and CNGC20 did not disrupt these cGMP-activated currents. The yellow fluorescent protein-CNGC5 and yellow fluorescent protein-CNGC6 proteins localize in the cell periphery. Cyclic AMP activated modest inward currents in both wild-type and cngc5cngc6 mutant guard cells. Moreover, cngc5 cngc6 double mutant guard cells exhibited functional abscisic acid (ABA)-activated hyperpolarization-dependent Ca2+-permeable cation channel currents, intact ABA-induced stomatal closing responses, and whole-plant stomatal conductance responses to darkness and changes in CO2 concentration. Furthermore, cGMP-activated currents remained intact in the growth controlled by abscisic acid2 and abscisic acid insensitive1 mutants. This research demonstrates that the CNGC5 and CNGC6 genes encode unique cGMP-activated nonselective Ca2+-permeable cation channels in the plasma membrane of Arabidopsis guard cells.Plants lose water via transpiration and take in CO2 for photosynthesis through stomatal pores. Each stomatal pore is surrounded by two guard cells, and stomatal movements are driven by the change of turgor pressure in guard cells. The intracellular second messenger Ca2+ functions in guard cell signal transduction (Schroeder and Hagiwara, 1989; McAinsh et al., 1990; Webb et al., 1996; Grabov and Blatt, 1998; Allen et al., 1999; MacRobbie, 2000; Mori et al., 2006; Young et al., 2006; Siegel et al., 2009; Chen et al., 2010; Hubbard et al., 2012). Plasma membrane ion channel activity and gene expression in guard cells are finely regulated by the intracellular free calcium concentration ([Ca2+]cyt; Schroeder and Hagiwara, 1989; Webb et al., 2001; Allen et al., 2002; Siegel et al., 2009; Kim et al., 2010; Stange et al., 2010). Ca2+-dependent protein kinases (CPKs) function as targets of the cytosolic Ca2+ signal, and several members of the CPK family have been shown to function in stimulus-induced stomatal closing, including the Arabidopsis (Arabidopsis thaliana) CPK3, CPK4, CPK6, CPK10, and CPK11 proteins (Mori et al., 2006; Zhu et al., 2007; Zou et al., 2010; Brandt et al., 2012; Hubbard et al., 2012). Further research found that several CPKs could activate the S-type anion channel SLAC1 in Xenopus laevis oocytes, including CPK21, CPK23, and CPK6 (Geiger et al., 2010; Brandt et al., 2012). At the same time, the Ca2+-independent protein kinase Open Stomata1 mediates stomatal closing and activates the S-type anion channel SLAC1 (Mustilli et al., 2002; Yoshida et al., 2002; Geiger et al., 2009; Lee et al., 2009; Xue et al., 2011), indicating that both Ca2+-dependent and Ca2+-independent pathways function in guard cells.Multiple essential factors of guard cell abscisic acid (ABA) signal transduction function in the regulation of Ca2+-permeable channels and [Ca2+]cyt elevations, including Abscisic Acid Insensitive1 (ABI1), ABI2, Enhanced Response to Abscisic Acid1 (ERA1), the NADPH oxidases AtrbohD and AtrbohF, the Guard Cell Hydrogen Peroxide-Resistant1 (GHR1) receptor kinase, as well as the Ca2+-activated CPK6 protein kinase (Pei et al., 1998; Allen et al., 1999, 2002; Kwak et al., 2003; Miao et al., 2006; Mori et al., 2006; Hua et al., 2012). [Ca2+]cyt increases result from both Ca2+ release from intracellular Ca2+ stores (McAinsh et al., 1992) and Ca2+ influx across the plasma membrane (Hamilton et al., 2000; Pei et al., 2000; Murata et al., 2001; Kwak et al., 2003; Hua et al., 2012). Electrophysiological analyses have characterized nonselective Ca2+-permeable channel activity in the plasma membrane of guard cells (Schroeder and Hagiwara, 1990; Hamilton et al., 2000; Pei et al., 2000; Murata et al., 2001; Köhler and Blatt, 2002; Miao et al., 2006; Mori et al., 2006; Suh et al., 2007; Vahisalu et al., 2008; Hua et al., 2012). However, the genetic identities of Ca2+-permeable channels in the plasma membrane of guard cells have remained unknown despite over two decades of research on these channel activities.The Arabidopsis genome includes 20 genes encoding cyclic nucleotide-gated channel (CNGC) homologs and 20 genes encoding homologs to animal Glu receptor channels (Lacombe et al., 2001; Kaplan et al., 2007; Ward et al., 2009), which have been proposed to function in plant cells as cation channels (Schuurink et al., 1998; Arazi et al., 1999; Köhler et al., 1999). Recent research has demonstrated functions of specific Glu receptor channels in mediating Ca2+ channel activity (Michard et al., 2011; Vincill et al., 2012). Previous studies have shown cAMP activation of nonselective cation currents in guard cells (Lemtiri-Chlieh and Berkowitz, 2004; Ali et al., 2007). However, only a few studies have shown the disappearance of a defined plasma membrane Ca2+ channel activity in plants upon mutation of candidate Ca2+ channel genes (Ali et al., 2007; Michard et al., 2011; Laohavisit et al., 2012; Vincill et al., 2012). Some CNGCs have been found to be involved in cation nutrient intake, including monovalent cation intake (Guo et al., 2010; Caballero et al., 2012), salt tolerance (Guo et al., 2008; Kugler et al., 2009), programmed cell death and pathogen responses (Clough et al., 2000; Balagué et al., 2003; Urquhart et al., 2007; Abdel-Hamid et al., 2013), thermal sensing (Finka et al., 2012; Gao et al., 2012), and pollen tube growth (Chang et al., 2007; Frietsch et al., 2007; Tunc-Ozdemir et al., 2013a, 2013b). Direct in vivo disappearance of Ca2+ channel activity in cngc disruption mutants has been demonstrated in only a few cases thus far (Ali et al., 2007; Gao et al., 2012). In this research, we show that CNGC5 and CNGC6 are required for a cyclic GMP (cGMP)-activated nonselective Ca2+-permeable cation channel activity in the plasma membrane of Arabidopsis guard cells.  相似文献   

4.
Guard cells use compensatory feedback controls to adapt to conditions that produce excessively open stomata.In the past 15 years or more, many mutants that are impaired in stimulus-induced stomatal closing and opening have been identified and functionally characterized in Arabidopsis (Arabidopsis thaliana), leading to a mechanistic understanding of the guard cell signal transduction network. However, evidence has only recently emerged that mutations impairing stomatal closure, in particular those in slow anion channel SLOW ANION CHANNEL-ASSOCIATED1 (SLAC1), unexpectedly also exhibit slowed stomatal opening responses. Results suggest that this compensatory slowing of stomatal opening can be attributed to a calcium-dependent posttranslational down-regulation of stomatal opening mechanisms, including down-regulation of inward K+ channel activity. Here, we discuss this newly emerging stomatal compensatory feedback control model mediated via constitutive enhancement (priming) of intracellular Ca2+ sensitivity of ion channel activity. The CALCIUM-DEPENDENT PROTEIN KINASE6 (CPK6) is strongly activated by physiological Ca2+ elevations and a model is discussed and open questions are raised for cross talk among Ca2+-dependent and Ca2+-independent guard cell signal transduction pathways and Ca2+ sensitivity priming mechanisms.Stomatal pores formed by two guard cells enable CO2 uptake from the atmosphere, but also ensure leaf cooling and provide a pulling force for nutrient uptake from the soil via transpiration. These vitally important processes are inevitably accompanied by water loss through stomata. Stomatal opening and closure is caused by the uptake and release of osmotically active substances and is tightly regulated by signaling pathways that lead to the activation or inactivation of guard cell ion channels and pumps. Potassium ions enter guard cells through the inward-rectifying K+ channels (K+in) during stomatal opening and are released via outward-rectifying K+ channels during stomatal closure (Schroeder et al., 1987; Hosy et al., 2003; Roelfsema and Hedrich 2005). Cytosolic Ca2+, an important second messenger in plants, mediates ion channel regulation, particularly down-regulation of inward-conducting K+in channels and activation of S-type anion channels, thus mediating stomatal closure and inhibiting stomatal opening (Schroeder and Hagiwara, 1989; Dodd et al., 2010; Kim et al., 2010). Stomatal closure is initiated by anion efflux via the slow S-type anion channel SLAC1 (Negi et al., 2008; Vahisalu et al., 2008; Kollist et al., 2011) and the voltage-dependent rapid R-type anion channel QUICK-ACTIVATING ANION CHANNEL1 (Meyer et al. 2010; Sasaki et al., 2010).In recent years, advances have been made toward understanding mechanisms mediating abscisic acid (ABA)-induced stomatal closure (Cutler et al., 2010; Kim et al., 2010; Raghavendra et al., 2010). The core ABA signaling module, consisting of PYR/RCAR (for pyrabactin resistance 1/regulatory components of ABA receptors) receptors, clade A protein phosphatases (PP2Cs), SNF-related protein kinase OPEN STOMATA1 (OST1), and downstream targets, is Ca2+-independent (Ma et al., 2009; Park et al., 2009; Hubbard et al., 2010). However, ABA-induced stomatal closure was reduced to only 30% of the normal stomatal closure response under conditions that inhibited intracellular cytosolic free calcium ([Ca2+]cyt) elevations in Arabidopsis (Siegel et al., 2009), consistent with previous findings in other plants (De Silva et al., 1985; Schwartz, 1985; McAinsh et al., 1991; MacRobbie, 2000). Together these and other studies show the importance of [Ca2+]cyt for a robust ABA-induced stomatal closure. Here, we discuss Ca2+-dependent and Ca2+-independent signaling pathways in guard cells and open questions on how these may work together.Plants carrying mutations in the SLAC1 anion channel have innately more open stomata, and exhibit clear impairments in ABA-, elevated CO2-, Ca2+-, ozone-, air humidity-, darkness-, and hydrogen peroxide-induced stomatal closure (Negi et al., 2008; Vahisalu et al., 2008; Merilo et al., 2013). Recent research, however, unexpectedly revealed that mutations in SLAC1 also down-regulate stomatal opening mechanisms and slow down stomatal opening (Laanemets et al., 2013).  相似文献   

5.
Ca2+-dependent protein kinases (CPKs) form a large family of 34 genes in Arabidopsis (Arabidopsis thaliana). Based on their dependence on Ca2+, CPKs can be sorted into three types: strictly Ca2+-dependent CPKs, Ca2+-stimulated CPKs (with a significant basal activity in the absence of Ca2+), and essentially calcium-insensitive CPKs. Here, we report on the third type of CPK, CPK13, which is expressed in guard cells but whose role is still unknown. We confirm the expression of CPK13 in Arabidopsis guard cells, and we show that its overexpression inhibits light-induced stomatal opening. We combine several approaches to identify a guard cell-expressed target. We provide evidence that CPK13 (1) specifically phosphorylates peptide arrays featuring Arabidopsis K+ Channel KAT2 and KAT1 polypeptides, (2) inhibits KAT2 and/or KAT1 when expressed in Xenopus laevis oocytes, and (3) closely interacts in plant cells with KAT2 channels (Förster resonance energy transfer-fluorescence lifetime imaging microscopy). We propose that CPK13 reduces stomatal aperture through its inhibition of the guard cell-expressed KAT2 and KAT1 channels.Stomata are microscopic organs at the leaf surface, each made of two so-called guard cells forming a pore. Opening or closing these pores is the way through which plants control their gas exchanges with the atmosphere (i.e. carbon dioxide uptake to feed the photosynthetic process and transpirational loss of water vapor). Stomatal movements result from osmotically driven fluxes of water, which follow massive exchanges of solutes, including K+ ions, between the guard cells and the surrounding tissues (Hetherington, 2001; Nilson and Assmann, 2007).Both Ca2+-dependent and Ca2+-independent signaling pathways are known to control stomatal movements (MacRobbie, 1993, 1998; Blatt, 2000; Webb et al., 2001; Mustilli et al., 2002; Israelsson et al., 2006; Marten et al., 2007; Laanemets et al., 2013). In particular, Ca2+ signals have been reported to promote stomatal closure through the inhibition of inward K+ channels and the activation of anion channels (Blatt, 1991, 1992, 2000; Thiel et al., 1992; Grabov and Blatt, 1999; Schroeder et al., 2001; Hetherington and Brownlee, 2004; Mori et al., 2006; Marten et al., 2007; Geiger et al., 2010; Brandt et al., 2012; Scherzer et al., 2012). However, little is known about the molecular identity of the links between Ca2+ events and Shaker K+ channel activity. Several kinases and phosphatases are believed to be involved in both the Ca2+-dependent and Ca2+-independent signaling pathways. Plants express two large kinase families whose activity is related to Ca2+ signaling. Firstly, CBL-interacting protein kinases (CIPKs; 25 genes in Arabidopsis [Arabidopsis thaliana]) are indirectly controlled by their interaction with a set of calcium sensors, the calcineurin B-like proteins (CBLs; 10 genes in Arabidopsis). This complex forms a fascinating network of potential Ca2+ signaling decoders (Luan, 2009; Weinl and Kudla, 2009), which have been addressed in numerous reports (Xu et al., 2006; Hu et al., 2009; Batistic et al., 2010; Held et al., 2011; Chen et al., 2013). In particular, some CBL-CIPK pairs have been shown to regulate Shaker channels such as Arabidopsis K+ Transporter1 (AKT1; Xu et al., 2006; Lan et al., 2011) or AKT2 (Held et al., 2011). Second, Ca2+-dependent protein kinases (CPKs) form an even larger family (34 genes in Arabidopsis) of proteins combining a kinase domain with the ability to bind Ca2+, thanks to the so-called EF hands (Harmon et al., 2000; Harper et al., 2004). CPKs, which, interestingly, are not found in animal cells, exhibit different calcium dependencies (Boudsocq et al., 2012). With respect to this, three types of CPKs can be considered: strictly Ca2+-dependent CPKs, Ca2+-stimulated CPKs (with a significant basal activity in the absence of Ca2+), and essentially Ca2+-insensitive CPKs (however, structurally close to kinases of groups 1 and 2).Pioneering work by Luan et al. (1993) demonstrated in Vicia faba guard cells that inward K+ channels were regulated by some Ca2+-dependent kinases. Then, such a Ca2+-dependent kinase was purified from guard cell protoplasts of V. faba and shown to actually phosphorylate the in vitro-translated KAT1 protein, a Shaker channel subunit natively expressed in Arabidopsis guard cells (Li et al., 1998). KAT1 regulation by CPK was shown by the inhibition of KAT1 currents after the coexpression of KAT1 and CDPK from soybean (Glycine max) in oocytes (Berkowitz et al., 2000). Since then, several cpk mutant lines of Arabidopsis have been shown to be impaired in stomatal movements, for example cpk10 (Ca2+ insensitive), cpk4/cpk11 (Ca2+ dependent), and cpk3/cpk6/cpk23 (Ca2+ dependent; Mori et al., 2006; Geiger et al., 2010; Munemasa et al., 2011; Hubbard et al., 2012).Of the nine genes encoding voltage-dependent K+ channels (Shaker) in Arabidopsis (Véry and Sentenac, 2002, 2003; Lebaudy et al., 2007; Hedrich, 2012), six are expressed in guard cells and play a role in stomatal movements: the Gated Outwardly-Rectifying K+ (GORK) gene, encoding an outward K+ channel subunit, and the AKT1, AKT2, Arabidopsis K+ Rectifying Channel1 (AtKC1), KAT1, and KAT2 genes, encoding inward K+ channel subunits (Pilot et al., 2001; Szyroki et al., 2001; Hosy et al., 2003; Pandey et al., 2007; Lebaudy et al., 2008a). Shaker channels result from the assembly of four subunits, and it has been shown that inward subunits tend to heterotetramerize, thus potentially widening the functional and regulatory scope of inward K+ conductance in guard cells (Xicluna et al., 2007; Jeanguenin et al., 2008; Lebaudy et al., 2008a, 2010). Inhibition of inward K+ channels has been shown to reduce stomatal opening (Liu et al., 2000; Kwak et al., 2001). This has grounded a strategy for disrupting inward K+ channel conductance in guard cells by expressing a nonfunctional KAT2 subunit (dominant negative mutation) in a kat2 knockout Arabidopsis line. The resulting Arabidopsis lines, named kincless, have no functional inward K+ channels and exhibit delayed stomatal opening (Lebaudy et al., 2008b) with, in the long term, a biomass reduction compared with the Arabidopsis wild-type line.Among the CPKs presumably expressed in Arabidopsis guard cells (Leonhardt et al., 2004), we looked for CPK13, which belongs to the atypical Ca2+-insensitive type of CPKs (Kanchiswamy et al., 2010; Boudsocq et al., 2012; Liese and Romeis, 2013) and whose role remains unknown in stomatal movements. Here, we confirm first that CPK13 kinase activity is independent of Ca2+ and show that CPK13 expression is predominant in Arabidopsis guard cells using CPK13-GUS lines. We then report that overexpression of CPK13 in Arabidopsis induces a dramatic default in stomatal aperture. Based on the previously reported kincless phenotype (Lebaudy et al., 2008b), we propose that CPK13 could reduce the activity of inward K+ channels in guard cells, particularly that of KAT2. We confirm this hypothesis by voltage-clamp experiments and show an inhibition of KAT2 and KAT1 activity by CPK13 (but not that of AKT2). In addition, we present peptide array phosphorylation assays showing that CPK13 targets, with some specificity, several KAT2 and KAT1 polypeptides. Finally, we demonstrate that KAT2 and CPK13 interact in planta using Förster resonance energy transfer (FRET)-fluorescence lifetime imaging microscopy (FLIM).  相似文献   

6.
Abscisic acid (ABA) induces stomatal closure and inhibits light-induced stomatal opening. The mechanisms in these two processes are not necessarily the same. It has been postulated that the ABA receptors involved in opening inhibition are different from those involved in closure induction. Here, we provide evidence that four recently identified ABA receptors (PYRABACTIN RESISTANCE1 [PYR1], PYRABACTIN RESISTANCE-LIKE1 [PYL1], PYL2, and PYL4) are not sufficient for opening inhibition in Arabidopsis (Arabidopsis thaliana). ABA-induced stomatal closure was impaired in the pyr1/pyl1/pyl2/pyl4 quadruple ABA receptor mutant. ABA inhibition of the opening of the mutant’s stomata remained intact. ABA did not induce either the production of reactive oxygen species and nitric oxide or the alkalization of the cytosol in the quadruple mutant, in accordance with the closure phenotype. Whole cell patch-clamp analysis of inward-rectifying K+ current in guard cells showed a partial inhibition by ABA, indicating that the ABA sensitivity of the mutant was not fully impaired. ABA substantially inhibited blue light-induced phosphorylation of H+-ATPase in guard cells in both the mutant and the wild type. On the other hand, in a knockout mutant of the SNF1-related protein kinase, srk2e, stomatal opening and closure, reactive oxygen species and nitric oxide production, cytosolic alkalization, inward-rectifying K+ current inactivation, and H+-ATPase phosphorylation were not sensitive to ABA.The phytohormone abscisic acid (ABA), which is synthesized in response to abiotic stresses, plays a key role in the drought hardiness of plants. Reducing transpirational water loss through stomatal pores is a major ABA response (Schroeder et al., 2001). ABA promotes the closure of open stomata and inhibits the opening of closed stomata. These effects are not simply the reverse of one another (Allen et al., 1999; Wang et al., 2001; Mishra et al., 2006).A class of receptors of ABA was identified (Ma et al., 2009; Park et al., 2009; Santiago et al., 2009; Nishimura et al., 2010). The sensitivity of stomata to ABA was strongly decreased in quadruple and sextuple mutants of the ABA receptor genes PYRABACTIN RESISTANCE/PYRABACTIN RESISTANCE-LIKE/REGULATORY COMPONENT OF ABSCISIC ACID RECEPTOR (PYR/PYL/RCAR; Nishimura et al., 2010; Gonzalez-Guzman et al., 2012). The PYR/PYL/RCAR receptors are involved in the early ABA signaling events, in which a sequence of interactions of the receptors with PROTEIN PHOSPHATASE 2Cs (PP2Cs) and subfamily 2 SNF1-RELATED PROTEIN KINASES (SnRK2s) leads to the activation of downstream ABA signaling targets in guard cells (Cutler et al., 2010; Kim et al., 2010; Weiner et al., 2010). Studies of Commelina communis and Vicia faba suggested that the ABA receptors involved in stomatal opening are not the same as the ABA receptors involved in stomatal closure (Allan et al., 1994; Anderson et al., 1994; Assmann, 1994; Schwartz et al., 1994). The roles of PYR/PYL/RCAR in either stomatal opening or closure remained to be elucidated.Blue light induces stomatal opening through the activation of plasma membrane H+-ATPase in guard cells that generates an inside-negative electrochemical gradient across the plasma membrane and drives K+ uptake through voltage-dependent inward-rectifying K+ channels (Assmann et al., 1985; Shimazaki et al., 1986; Blatt, 1987; Schroeder et al., 1987; Thiel et al., 1992). Phosphorylation of the penultimate Thr of the plasma membrane H+-ATPase is a prerequisite for blue light-induced activation of the H+-ATPase (Kinoshita and Shimazaki, 1999, 2002). ABA inhibits H+-ATPase activity through dephosphorylation of the penultimate Thr in the C terminus of the H+-ATPase in guard cells, resulting in prevention of the opening (Goh et al., 1996; Zhang et al., 2004; Hayashi et al., 2011). Inward-rectifying K+ currents (IKin) of guard cells are negatively regulated by ABA in addition to through the decline of the H+ pump-driven membrane potential difference (Schroeder and Hagiwara, 1989; Blatt, 1990; McAinsh et al., 1990; Schwartz et al., 1994; Grabov and Blatt, 1999; Saito et al., 2008). This down-regulation of ion transporters by ABA is essential for the inhibition of stomatal opening.A series of second messengers has been shown to mediate ABA-induced stomatal closure. Reactive oxygen species (ROS) produced by NADPH oxidases play a crucial role in ABA signaling in guard cells (Pei et al., 2000; Zhang et al., 2001; Kwak et al., 2003; Sirichandra et al., 2009; Jannat et al., 2011). Nitric oxide (NO) is an essential signaling component in ABA-induced stomatal closure (Desikan et al., 2002; Guo et al., 2003; Garcia-Mata and Lamattina, 2007; Neill et al., 2008). Alkalization of cytosolic pH in guard cells is postulated to mediate ABA-induced stomatal closure in Arabidopsis (Arabidopsis thaliana) and Pisum sativum and Paphiopedilum species (Irving et al., 1992; Gehring et al., 1997; Grabov and Blatt, 1997; Suhita et al., 2004; Gonugunta et al., 2008). These second messengers transduce environmental signals to ion channels and ion transporters that create the driving force for stomatal movements (Ward et al., 1995; MacRobbie, 1998; Garcia-Mata et al., 2003).In this study, we examined the mobilization of second messengers, the inactivation of IKin, and the suppression of H+-ATPase phosphorylation evoked by ABA in Arabidopsis mutants to clarify the downstream signaling events of ABA signaling in guard cells. The mutants included a quadruple mutant of PYR/PYL/RCARs, pyr1/pyl1/pyl2/pyl4, and a mutant of a SnRK2 kinase, srk2e.  相似文献   

7.
8.
Yeast elicitor (YEL) induces stomatal closure that is mediated by a Ca2+-dependent signaling pathway. A Ca2+-dependent protein kinase, CPK6, positively regulates activation of ion channels in abscisic acid and methyl jasmonate signaling, leading to stomatal closure in Arabidopsis (Arabidopsis thaliana). YEL also inhibits light-induced stomatal opening. However, it remains unknown whether CPK6 is involved in induction by YEL of stomatal closure or in inhibition by YEL of light-induced stomatal opening. In this study, we investigated the roles of CPK6 in induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening in Arabidopsis. Disruption of CPK6 gene impaired induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening. Activation by YEL of nonselective Ca2+-permeable cation channels was impaired in cpk6-2 guard cells, and transient elevations elicited by YEL in cytosolic-free Ca2+ concentration were suppressed in cpk6-2 and cpk6-1 guard cells. YEL activated slow anion channels in wild-type guard cells but not in cpk6-2 or cpk6-1 and inhibited inward-rectifying K+ channels in wild-type guard cells but not in cpk6-2 or cpk6-1. The cpk6-2 and cpk6-1 mutations inhibited YEL-induced hydrogen peroxide accumulation in guard cells and apoplast of rosette leaves but did not affect YEL-induced hydrogen peroxide production in the apoplast of rosette leaves. These results suggest that CPK6 positively functions in induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening in Arabidopsis and is a convergent point of signaling pathways for stomatal closure in response to abiotic and biotic stress.Stomata, formed by pairs of guard cells, play a critical role in regulation of plant CO2 uptake and water loss, thus critically influencing plant growth and water stress responsiveness. Guard cells respond to a variety of abiotic and biotic stimuli, such as light, drought, and pathogen attack (Israelsson et al., 2006; Shimazaki et al., 2007; Melotto et al., 2008).Elicitors derived from microbial surface mimic pathogen attack and induce stomatal closure in various plant species such as Solanum lycopersicum (Lee et al., 1999), Commelina communis (Lee et al., 1999), Hordeum vulgare (Koers et al., 2011), and Arabidopsis (Arabidopsis thaliana; Melotto et al., 2006; Khokon et al., 2010). Yeast elicitor (YEL) induces stomatal closure in Arabidopsis (Klüsener et al., 2002; Khokon et al., 2010; Salam et al., 2013). Our recent studies showed that YEL inhibits light-induced stomatal opening and that protein phosphorylation is involved in induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening (Salam et al., 2013).Cytosolic Ca2+ has long been recognized as a conserved second messenger in stomatal movement (Shimazaki et al., 2007; Roelfsema and Hedrich 2010; Hubbard et al., 2012). Elevation of cytosolic free Ca2+ concentration ([Ca2+]cyt) is triggered by influx of Ca2+ from apoplast and release of Ca2+ from intracellular stores in guard cell signaling (Leckie et al., 1998; Hamilton et al., 2000; Pei et al., 2000; Garcia-Mata et al., 2003; Lemtiri-Chlieh et al., 2003). The influx of Ca2+ is carried by nonselective Ca2+-permeable cation (ICa) channels that are activated by plasma membrane hyperpolarization and H2O2 (Pei et al., 2000; Murata et al., 2001; Kwak et al., 2003). Elevation of [Ca2+]cyt activates slow anion (S-type) channels and down-regulates inward-rectifying potassium (Kin) channels in guard cells (Schroeder and Hagiwara, 1989; Grabov and Blatt, 1999). The activation of S-type channels is a hallmark of stomatal closure, and the suppression of Kin channels is favorable to stomatal closure but not to stomatal opening (Pei et al., 1997; Kwak et al., 2001; Xue et al., 2011; Uraji et al., 2012).YEL induces stomatal closure with extracellular H2O2 production, intracellular H2O2 accumulation, activation of ICa channels, and transient [Ca2+]cyt elevations (Klüsener et al., 2002; Khokon et al., 2010). However, it remains to be clarified whether YEL activates S-type channels and inhibits Kin channels in guard cells.Calcium-dependent protein kinases (CDPKs) are regulators in Ca2+-dependent guard cell signaling (Mori et al., 2006; Zhu et al., 2007; Geiger et al., 2010, 2011; Zou et al., 2010; Munemasa et al., 2011; Brandt et al., 2012; Scherzer et al., 2012). In guard cells, CDPKs regulate activation of S-type and ICa channels and inhibition of Kin channels (Mori et al., 2006; Zou et al., 2010; Munemasa et al., 2011). A CDPK, CPK6, positively regulates activation of S-type channels and ICa channels without affecting H2O2 production in abscisic acid (ABA)- and methyl jasmonate (MeJA)-induced stomatal closure (Mori et al., 2006; Munemasa et al., 2011). CPK6 phosphorylates and activates SLOW ANION CHANNEL-ASSOCIATED1 expressed in Xenopus spp. oocyte (Brandt et al., 2012; Scherzer et al., 2012). These findings underline the role of CPK6 in regulation of ion channel activation and stomatal movement, leading us to test whether CPK6 regulates the induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening.In this study, we investigated activation of S-type channels and inhibition of Kin channels by YEL and roles of CPK6 in induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening. For this purpose, we examined the effects of mutation of CPK6 on induction by YEL of stomatal closure and inhibition by YEL of light-induced stomatal opening, activation of ICa channels, transient [Ca2+]cyt elevations, activation of S-type channels, inhibition of Kin channels, H2O2 production in leaves, and H2O2 accumulation in leaves and guard cells.  相似文献   

9.
10.
Stomatal transpiration is at the center of a crisis in water availability and crop production that is expected to unfold over the next 20 to 30 years. Global water usage has increased 6-fold in the past 100 years, twice as fast as the human population, and is expected to double again before 2030, driven mainly by irrigation and agriculture. Guard cell membrane transport is integral to controlling stomatal aperture and offers important targets for genetic manipulation to improve crop performance. However, its complexity presents a formidable barrier to exploring such possibilities. With few exceptions, mutations that increase water use efficiency commonly have been found to do so with substantial costs to the rate of carbon assimilation, reflecting the trade-off in CO2 availability with suppressed stomatal transpiration. One approach yet to be explored in detail relies on quantitative systems analysis of the guard cell. Our deep knowledge of transport and homeostasis in these cells gives real substance to the prospect for reverse engineering of stomatal responses, using in silico design in directing genetic manipulation for improved water use and crop yields. Here we address this problem with a focus on stomatal kinetics, taking advantage of the OnGuard software and models of the stomatal guard cell recently developed for exploring stomatal physiology. Our analysis suggests that manipulations of single transporter populations are likely to have unforeseen consequences. Channel gating, especially of the dominant K+ channels, appears the most favorable target for experimental manipulation.Stomata are pores that provide the major route for gaseous exchange across the impermeable cuticle of leaves and stems (Hetherington and Woodward, 2003). They open and close in response to exogenous and endogenous signals and thereby control the exchange of gases, most importantly water vapor and CO2, between the interior of the leaf and the atmosphere. Stomata exert major controls on the water and carbon cycles of the world (Schimel et al., 2001) and can limit photosynthetic rates by 50% or more when demand exceeds water supply (Ni and Pallardy, 1992). Stomatal transpiration is at the center of a crisis in water availability and crop production that is expected to unfold over the next 20 to 30 years; indeed, global water usage has increased 6-fold in the past 100 years, twice as fast as the human population, and is expected to double again before 2030, driven mainly by irrigation and agriculture (United Nations Educational, Scientific and Cultural Organization, 2009).Guard cell transport is integral to controlling stomatal aperture. Guard cells surround the stomatal pore and respond in a well-defined manner to an array of extracellular signals, including light, to regulate its aperture. Guard cells coordinate membrane transport within a complex network of intracellular signals (Willmer and Fricker, 1996; Blatt, 2000a, 2000b; Hetherington and Woodward, 2003; Shimazaki et al., 2007) to regulate fluxes, mainly of K+, Cl, and malate, driving cell turgor and stomatal aperture. Our deep knowledge of these processes has made the guard cell the best known of plant cell models for membrane transport, signaling, and homeostasis (Willmer and Fricker, 1996; Blatt, 2000b; Roelfsema and Hedrich, 2010; Hills et al., 2012). This knowledge gives real substance to the prospect for reverse engineering of stomatal responses, using in silico design in directing genetic manipulation for improved crop yields, especially under water-limited conditions.Water use efficiency (WUE; defined as the amount of dry matter produced per unit of water transpired) is directly related to stomatal function. Thus, at the practical level, stomata represent an important target for breeders interested in manipulating crop performance. A large body of data relates stomata, transpiration, and carbon assimilation (Willmer and Fricker, 1996; Farquhar et al., 2001; Hetherington and Woodward, 2003; Lawson et al., 2011). Several examples illustrate how manipulating of stomatal characteristics can affect WUE (Fischer et al., 1998; Rebetzke et al., 2002; Masle et al., 2005; Eisenach et al., 2012). With few exceptions, however, mutations that increase WUE commonly do so at the expense of carbon assimilation, reflecting the trade-off in CO2 availability with suppressed stomatal transpiration.Stomatal movements generally lag behind short-term changes in available light associated with sunflecks and shadeflecks (Pearcy, 1990; Lawson et al., 2012; Lawson and Blatt, 2014). This hysteresis in response, between stomatal aperture and gas exchange on one hand and photosynthetic capacity on the other, can lead alternately to periods of assimilation limited by stomatal conductance, and of high transpiration without corresponding rates of assimilation (Lawson et al., 2011). It has been argued that such hysteresis in stomatal responsiveness with the demand for CO2 erodes assimilation and WUE, with substantial consequences for long-term yield (Vico et al., 2011; Eisenach et al., 2012; Lawson et al., 2012; Lawson and Blatt, 2014). If so, then improving WUE with gains in assimilation should be possible if the speed of stomatal responsiveness can be enhanced. However, the complexity of guard cell transport presents a formidable barrier to exploring such possibilities. Here we address this problem, taking advantage of OnGuard models of the stomatal guard cell. We explore in silico the potential for enhancing stomatal kinetics through single transporter (single gene product) manipulations. Our results identify the gating of the dominant K+ channels as the most promising target for experimental manipulation.  相似文献   

11.
Stomatal responses to vapor pressure deficit (VPD) are a principal means by which vascular land plants regulate daytime transpiration. While much work has focused on characterizing and modeling this response, there remains no consensus as to the mechanism that drives it. Explanations range from passive regulation by leaf hydration to biochemical regulation by the phytohormone abscisic acid (ABA). We monitored ABA levels, leaf gas exchange, and water status in a diversity of vascular land plants exposed to a symmetrical, mild transition in VPD. The stomata in basal lineages of vascular plants, including gymnosperms, appeared to respond passively to changes in leaf water status induced by VPD perturbation, with minimal changes in foliar ABA levels and no hysteresis in stomatal action. In contrast, foliar ABA appeared to drive the stomatal response to VPD in our angiosperm samples. Increased foliar ABA level at high VPD in angiosperm species resulted in hysteresis in the recovery of stomatal conductance; this was most pronounced in herbaceous species. Increased levels of ABA in the leaf epidermis were found to originate from sites of synthesis in other parts of the leaf rather than from the guard cells themselves. The transition from a passive regulation to ABA regulation of the stomatal response to VPD in the earliest angiosperms is likely to have had critical implications for the ecological success of this lineage.Plants continuously regulate transpiration by controlling the aperture of the stomatal pores on the surface of the leaf. The principal atmospheric determinant of stomatal aperture is the humidity of the air, which can be expressed as the vapor pressure difference between the leaf and the atmosphere. Stomatal responses to atmospheric vapor pressure deficit (VPD) have been well characterized across the diversity of vascular plant species (Darwin, 1898; Lange et al., 1971; Turner et al., 1984; Franks and Farquhar, 1999; Oren et al., 1999; Brodribb and McAdam, 2011; Mott and Peak, 2013), with stomata typically closing at high VPD and opening at low VPD. This comprehensive characterization has allowed for the development of highly effective empirical and mechanistic models of leaf gas exchange that provide robust predictions of the responses of transpiration to changes in VPD (Buckley et al., 2003; Katul et al., 2009; Damour et al., 2010; Medlyn et al., 2011). Despite the success of this modeling, the mechanism for the stomatal response to VPD remains poorly understood (Damour et al., 2010). Different hypotheses range from one extreme, whereby stomata respond passively through changes in leaf water content induced by the VPD or humidity perturbation (Lange et al., 1971; Mott and Peak, 2013), to the other extreme, whereby stomata close uniquely in response to the phytohormone abscisic acid (ABA; Xie et al., 2006; Bauer et al., 2013).From the earliest recognition that stomata open and close by changes in guard cell turgor (Heath, 1938), there have been many attempts to link the passive changes in water status that occur during VPD or humidity transitions with stomatal responses to VPD or humidity (Lange et al., 1971; Mott and Peak, 2013). Studies have suggested that changes in atmospheric water content passively drive stomatal responses by changing bulk leaf water status, which in turn changes guard cell turgor (Oren et al., 1999), or alternatively by changing guard cell turgor directly (Mott and Peak, 2013). Models based on these entirely passive processes are highly effective in predicting steady-state stomatal conductance (gs) in response to changes in VPD or humidity in angiosperms (Mott and Peak, 2013).While hydraulic models provide robust predictions of steady-state gs, they are less effective at predicting the dynamic responses of stomata to short-term perturbations, particularly with respect to the wrong-way responses that typically occur as transients (Buckley, 2005), as well as feed-forward behavior (Farquhar, 1978; Bunce, 1997; Franks et al., 1997; Tardieu and Simonneau, 1998; Ocheltree et al., 2014; compare with Mott and Peak, 2013). Although some of these models provide a pathway for incorporating the effect of ABA (Buckley, 2005), a lack of knowledge of ABA dynamics or action makes it difficult to integrate the influence of this active regulator of guard cell aperture into models. The stomatal behavior of single gene mutants (most notably the ABA synthesis and signaling mutants of Arabidopsis) strongly supports a role for ABA in mediating standard stomatal responses to changes in VPD. The stomata of these mutants are known to have less pronounced responses to a reduction in relative humidity compared with wild-type plants (Xie et al., 2006). Recently, molecular work has shown that guard cells express many of the genes required to synthesize ABA (Okamoto et al., 2009; Bauer et al., 2013), with molecular proxies for ABA level also indicating that the biochemical activity of ABA in the guard cell may increase following short-term exposure of leaves to a reduction in relative humidity (Waadt et al., 2014). These findings suggest a role for ABA in regulating stomatal responses to VPD and have led some to the conclusion that ABA synthesized autonomously by the guard cells is the predominant mechanism for stomatal responses to increased VPD (Bauer et al., 2013).Although the experimental evidence from molecular studies presents an argument for the role of ABA in the responses of stomata to changes in VPD, very few studies have quantified changes in ABA level in response to VPD. It is well established that ABA levels in leaves and guard cells can increase following the imposition of turgor loss or water stress (Pierce and Raschke, 1980; Harris et al., 1988; Harris and Outlaw, 1991). However, only a few studies have reported increases in foliar ABA level in response to high VPD (Bauerle et al., 2004; Giday et al., 2013), and none have investigated whether these observed dynamic changes or differences in ABA level were functionally relevant for stomatal control. In addition, no study has quantified the levels of ABA in guard cells during a transition in VPD.Here, we investigate the relative importance of ABA for the stomatal response to VPD in whole plants, sampled from across the vascular land plant lineage. We provide, to our knowledge, the first functional assessment of changes in ABA levels driving stomatal responses to VPD as well as critically investigate the recent suggestion that stomatal responses to VPD are driven by an autonomous guard cell synthesis of ABA.  相似文献   

12.
13.
14.
The requirements of the water transport system of small herbaceous species differ considerably from those of woody species. Despite their ecological importance for many biomes, knowledge regarding herb hydraulics remains very limited. We compared key hydraulic features (vulnerability to drought-induced hydraulic decline, pressure-volume relations, onset of cellular damage, in situ variation of water potential, and stomatal conductance) of three Ranunculus species differing in their soil humidity preferences and ecological amplitude. All species were very vulnerable to water stress (50% reduction in whole-leaf hydraulic conductance [kleaf] at −0.2 to −0.8 MPa). In species with narrow ecological amplitude, the drought-exposed Ranunculus bulbosus was less vulnerable to desiccation (analyzed via loss of kleaf and turgor loss point) than the humid-habitat Ranunculus lanuginosus. Accordingly, water stress-exposed plants from the broad-amplitude Ranunculus acris revealed tendencies toward lower vulnerability to water stress (e.g. osmotic potential at full turgor, cell damage, and stomatal closure) than conspecific plants from the humid site. We show that small herbs can adjust to their habitat conditions on interspecific and intraspecific levels in various hydraulic parameters. The coordination of hydraulic thresholds (50% and 88% loss of kleaf, turgor loss point, and minimum in situ water potential) enabled the study species to avoid hydraulic failure and damage to living cells. Reversible recovery of hydraulic conductance, desiccation-tolerant seeds, or rhizomes may allow them to prioritize toward a more efficient but vulnerable water transport system while avoiding the severe effects that water stress poses on woody species.The resistance of terrestrial plant species to water stress is an important determinant of their spatial distribution (Engelbrecht et al., 2007), and in a globally changing climate, resistance to water stress-induced damage to the water transport system is an essential parameter for the prediction of species survival and future changes in biodiversity and species composition. The hydraulic vulnerability to drought-induced embolism in woody species is adjusted to their environmental conditions on a global scale (Choat et al., 2012), but increased hydraulic safety entails several tradeoffs, such as transport efficiency, as plants can reduce the risk of hydraulic failure by the formation of smaller and hydraulically less efficient conduits (Tyree et al., 1994; Wagner et al., 1998, Gleason et al., 2016).While many studies have dealt with the hydraulics of woody species (Maherali et al., 2004; Choat et al., 2012), the hydraulic characteristics of small herbaceous species are largely understudied, and our knowledge regarding their water transport system, including vulnerability to drought-induced loss of conductivity, remains very limited (Kocacinar and Sage, 2003; Brodribb and Holbrook, 2004; Iovi et al., 2009; Holloway-Phillips and Brodribb, 2011a; Tixier et al., 2013). Yet, herbs play an important ecological role in many biomes, such as grasslands and the alpine zone (Billings and Mooney, 1968; Gilliam, 2007; Scholz et al., 2010), and represent a significant percentage of crop plants (Monfreda et al., 2008); therefore, they are interesting from an ecophysiological viewpoint.The requirements of the water transport system of herbs differ in several ways from those of woody species (Mencuccini, 2003). First, transport distances and supported leaf areas are much smaller in herbs, and reversible extraxylary limitations (Brodribb and Holbrook, 2005; Kim and Steudle, 2007) may have a stronger effect on hydraulic conductivity. Second, herbs often do not feature an essential, lignified main axis designed for long-term functioning, in which hydraulic failure may mean whole-plant mortality. Third, small plants may more easily restore water transport capacity in the case of embolism formation, because, due to smaller size, a water potential (Ψ) at or near zero can be reached quickly under favorable conditions due to positive root pressure (Tyree and Sperry, 1989; Tyree and Ewers, 1991; Stiller and Sperry, 2002; Tyree and Zimmermann, 2002; Ganthaler and Mayr, 2015). Thus, herbs may not be as threatened by drought-induced failure as taller woody species.Hydraulic failure occurs when vulnerability thresholds are exceeded under water stress and the xylem tension is suddenly released as water is replaced by air, effectively blocking water transport in the affected conduits (Tyree and Zimmermann, 2002). Vulnerability to drought-induced embolism can thus be analyzed hydraulically by measuring decreases in conductance during increasing water stress (Sperry et al., 1988a; Cochard, 2002; Cochard et al., 2013), by noninvasive imaging of xylem water content (Choat et al., 2010; Brodersen et al., 2013; Cochard et al., 2015), or indirectly by monitoring of ultrasonic acoustic emissions (UE) which occur during the spontaneous energy release at embolism formation (Milburn and Johnson, 1966; Mayr and Rosner, 2011; Ponomarenko et al., 2014). However, the usability of UE analysis to noninvasively study hydraulic vulnerability remains under discussion (Sandford and Grace, 1985; Kikuta, 2003; Wolkerstorfer et al., 2012; Cochard et al., 2013; Rosner, 2015).On the leaf level, desiccation may induce turgor loss, decreases in mesophyll conductance (Kim and Steudle, 2007; Scoffoni et al., 2014), or cell collapse (Blackman et al., 2010; Holloway-Phillips and Brodribb, 2011a), and limit hydraulic conductance, gas exchange, and thus photosynthesis before embolism occurs. When water stress increases further, living cells may incur damage resulting in tissue and leaf mortality.In this study, we compared key hydraulic features (vulnerability to drought-induced loss of conductance as measured via hydraulic flow and xylem staining, pressure-volume relations, and the onset of cellular damage) of three species from the genus Ranunculus, which is distributed almost worldwide and contains species adapted to dry and humid habitats as well as species with broad ecological amplitude. We hypothesized that herbaceous species would be more vulnerable to water stress than woody species but also would show interspecific and intraspecific adjustments in hydraulic vulnerability based on the water availability of their respective habitats. Namely, the hydraulics of species with narrow ecological amplitude were hypothesized to reflect their respective habitat conditions, and broad-amplitude species should show adequate intraspecific variation to enable growth in dry and humid habitats.  相似文献   

15.
16.
The FLAGELLIN-SENSING2 (FLS2) receptor kinase recognizes bacterial flagellin and initiates a battery of downstream defense responses to reduce bacterial invasion through stomata in the epidermis and bacterial multiplication in the apoplast of infected plants. Recent studies have shown that during Pseudomonas syringae pv tomato (Pst) DC3000 infection of Arabidopsis (Arabidopsis thaliana), FLS2-mediated immunity is actively suppressed by effector proteins (such as AvrPto and AvrPtoB) secreted through the bacterial type III secretion system (T3SS). We provide evidence here that T3SS effector-based suppression does not appear to be sufficient to overcome FLS2-based immunity during Pst DC3000 infection, but that the phytotoxin coronatine (COR) produced by Pst DC3000 also plays a critical role. COR-deficient mutants of Pst DC3000 are severely reduced in virulence when inoculated onto the leaf surface of wild-type Columbia-0 plants, but this defect was rescued almost fully in fls2 mutant plants. Although bacteria are thought to carry multiple microbe-associated molecular patterns, stomata of fls2 plants are completely unresponsive to COR-deficient mutant Pst DC3000 bacteria. The responses of fls2 plants were similar to those of the Arabidopsis G-protein alpha subunit1-3 mutant, which is defective in abscisic acid-regulated stomatal closure, but were distinct from those of the Arabidopsis non-expressor of PR genes1 mutant, which is defective in salicylic acid-dependent stomatal closure and apoplast defense. Epistasis analyses show that salicylic acid signaling acts upstream of abscisic acid signaling in bacterium-triggered stomatal closure. Taken together, these results suggest a particularly important role of FLS2-mediated resistance to COR-deficient mutant Pst DC3000 bacteria, and nonredundant roles of COR and T3SS effector proteins in the suppression of FLS2-mediated resistance in the Arabidopsis-Pst DC3000 interaction.Stomata are microscopic pores formed by pairs of guard cells in the epidermis of terrestrial plants; they are essential for CO2 and water exchange with the environment. Plants regulate the stomatal aperture in response to changing abiotic environmental conditions (e.g. light, humidity, CO2 concentration) to optimize CO2 uptake and water transpiration. The molecular mechanisms underlying the stomatal regulation in response to abiotic signals are a subject of intense studies. Research in this area has uncovered many signaling components, indicating that stomatal guard cells have one of the most dynamic regulatory networks in plants (Schroeder et al., 2001; Shimazaki et al., 2007; Neill et al., 2008; Wang and Song, 2008).Stomatal openings are also a major route of pathogen entry into the plant (Melotto et al., 2006). Accordingly, guard cells have developed mechanisms to regulate stomatal aperture in response to pathogens. Melotto and colleagues found that the bacterial pathogen Pseudomonas syringae pv tomato (Pst) strain DC3000 induces stomatal closure in Arabidopsis (Arabidopsis thaliana) within 1 h post inoculation. However, after 3 to 4 h, stomata reopen (Melotto et al., 2006). The ability of Pst DC3000 to reopen stomata is dependent on the polyketide toxin coronatine (COR), a virulence factor that had previously been shown to be important for bacterial multiplication within the mesophyll space, disease symptom development, and induction of systemic susceptibility of infected plants (Mittal and Davis, 1995; Bender et al., 1999; Budde and Ullrich, 2000; Brooks et al., 2004; Cui et al., 2005; Melotto et al., 2008b). Stomatal reopening by Pst DC3000 was also shown to be dependent on the RPM1-INTERACTING PROTEIN4 in Arabidopsis (Liu et al., 2009). Recently, another bacterial pathogen, Xanthomonas campestris pv campestris, was shown to cause stomatal closure and subsequent reopening during infection (Gudesblat et al., 2009). In this case, a virulence factor of smaller than 2 kD was identified, but the molecular identity of this virulence factor is not yet known. In fungal pathogens, examples of virulence factors that inhibit stomatal closure include fusicoccin (Turner and Graniti, 1969; Assmann and Schwartz, 1992; Kinoshita and Shimazaki, 2001) and oxalic acid (Guimaraes and Stotz, 2004), although their role in pathogen invasion has not been established.Stomatal guard cells also respond to purified microbe-associated molecular patterns (MAMPs), such as chitosan, a polymer of β-1,4-glucosamine residues derived from fungal chitin (Lee et al., 1999; Amborabe et al., 2008), flg22, a 22-amino acid peptide derived from bacterial flagellin (Melotto et al., 2006; Cho et al., 2008; Desikan et al., 2008; Zhang et al., 2008), and bacterial lipopolysaccharides (LPSs; Melotto et al., 2006; Cho et al., 2008). Peptidoglycan, derived from Gram-positive bacteria, is shown to be able to induce plant innate immune responses (Gust et al., 2007; Erbs et al., 2008). However, peptidoglycan has not yet been shown to trigger stomatal responses. MAMPs are recognized by plant pattern-recognition receptors, such as Arabidopsis proteins FLAGELLIN-SENSING2 (FLS2) that recognizes bacterial flagellin (Gómez-Gómez and Boller, 2000), EF-TU RECEPTOR (EFR) that recognizes bacterial elongation factor TU (Zipfel et al., 2006), and CHITIN ELICITOR RECEPTOR KINASE1 (CERK1) that perceives an unknown MAMP from Pst DC3000 (Gimenez-Ibanez et al., 2009a, 2009b). In the case of flg22-induced stomatal closure, FLS2 is required (Melotto et al., 2006). Stomata from fls2 mutant plants, however, still respond to purified LPS (Melotto et al., 2006), illustrating both specificity in MAMP recognition by guard cells and the capacity of guard cells to recognize multiple MAMPs (Melotto et al., 2006). However, it has not been formally proven that the perception of any individual MAMPs actually contributes to stomatal closure induced by live bacteria, as bacteria produce many other factors in the context of an infection.Studies using purified MAMPs have shown that stomatal closure in response to biotic signals requires the phytohormone abscisic acid (ABA), the guard cell-specific OPEN STOMATA1 (OST1) kinase, the production of reactive oxygen species and nitric oxide, the heterotrimeric G protein, and the regulation of K+ channels—all of which are hallmarks of abiotic signal-induced stomatal closure (Melotto et al., 2006; Neill et al., 2008; Zhang et al., 2008). These findings suggest that the guard cell signal transductions in response to biotic and abiotic signals share common steps. Besides shared signaling components, however, MAMP-triggered stomatal closure also requires the plant defense hormone salicylic acid (SA; Melotto et al., 2006). At present, it is not clear whether SA per se or a downstream signaling component, such as the NON-EXPRESSOR OF PR GENES1 (NPR1), is required for stomatal closure. Nor do we understand the epistatic relationship between SA and ABA signaling in the regulation of bacterium/MAMP-triggered stomatal closure.In this study, we conducted experiments to further characterize stomatal regulation during Pst DC3000 infection of Arabidopsis plants. In particular, we sought to determine (1) whether the perception of well-documented MAMPs indeed contributes to stomatal closure in response to live bacteria, (2) the roles of the heterotrimeric G protein (involved in ABA signaling) and NPR1 (involved in SA signaling) in stomatal response during bacterial infection, and (3) the relationship between SA signaling and ABA signaling in regulating bacterium-triggered stomatal closure. These experiments revealed a critical role of FLS2 in mediating disease resistance against COR-deficient mutant Pst DC3000 bacteria.  相似文献   

17.
The role of calcium-mediated signaling has been extensively studied in plant responses to abiotic stress signals. Calcineurin B-like proteins (CBLs) and CBL-interacting protein kinases (CIPKs) constitute a complex signaling network acting in diverse plant stress responses. Osmotic stress imposed by soil salinity and drought is a major abiotic stress that impedes plant growth and development and involves calcium-signaling processes. In this study, we report the functional analysis of CIPK21, an Arabidopsis (Arabidopsis thaliana) CBL-interacting protein kinase, ubiquitously expressed in plant tissues and up-regulated under multiple abiotic stress conditions. The growth of a loss-of-function mutant of CIPK21, cipk21, was hypersensitive to high salt and osmotic stress conditions. The calcium sensors CBL2 and CBL3 were found to physically interact with CIPK21 and target this kinase to the tonoplast. Moreover, preferential localization of CIPK21 to the tonoplast was detected under salt stress condition when coexpressed with CBL2 or CBL3. These findings suggest that CIPK21 mediates responses to salt stress condition in Arabidopsis, at least in part, by regulating ion and water homeostasis across the vacuolar membranes.Drought and salinity cause osmotic stress in plants and severely affect crop productivity throughout the world. Plants respond to osmotic stress by changing a number of cellular processes (Xiong et al., 1999; Xiong and Zhu, 2002; Bartels and Sunkar, 2005; Boudsocq and Lauriére, 2005). Some of these changes include activation of stress-responsive genes, regulation of membrane transport at both plasma membrane (PM) and vacuolar membrane (tonoplast) to maintain water and ionic homeostasis, and metabolic changes to produce compatible osmolytes such as Pro (Stewart and Lee, 1974; Krasensky and Jonak, 2012). It has been well established that a specific calcium (Ca2+) signature is generated in response to a particular environmental stimulus (Trewavas and Malhó, 1998; Scrase-Field and Knight, 2003; Luan, 2009; Kudla et al., 2010). The Ca2+ changes are primarily perceived by several Ca2+ sensors such as calmodulin (Reddy, 2001; Luan et al., 2002), Ca2+-dependent protein kinases (Harper and Harmon, 2005), calcineurin B-like proteins (CBLs; Luan et al., 2002; Batistič and Kudla, 2004; Pandey, 2008; Luan, 2009; Sanyal et al., 2015), and other Ca2+-binding proteins (Reddy, 2001; Shao et al., 2008) to initiate various cellular responses.Plant CBL-type Ca2+ sensors interact with and activate CBL-interacting protein kinases (CIPKs) that phosphorylate downstream components to transduce Ca2+ signals (Liu et al., 2000; Luan et al., 2002; Batistič and Kudla, 2004; Luan, 2009). In several plant species, multiple members have been identified in the CBL and CIPK family (Luan et al., 2002; Kolukisaoglu et al., 2004; Pandey, 2008; Batistič and Kudla, 2009; Weinl and Kudla, 2009; Pandey et al., 2014). Involvement of specific CBL-CIPK pair to decode a particular type of signal entails the alternative and selective complex formation leading to stimulus-response coupling (D’Angelo et al., 2006; Batistič et al., 2010).Several CBL and CIPK family members have been implicated in plant responses to drought, salinity, and osmotic stress based on genetic analysis of Arabidopsis (Arabidopsis thaliana) mutants (Zhu, 2002; Cheong et al., 2003, 2007; Kim et al., 2003; Pandey et al., 2004, 2008; D’Angelo et al., 2006; Qin et al., 2008; Tripathi et al., 2009; Held et al., 2011; Tang et al., 2012; Drerup et al., 2013; Eckert et al., 2014). A few CIPKs have also been functionally characterized by gain-of-function approach in crop plants such as rice (Oryza sativa), pea (Pisum sativum), and maize (Zea mays) and were found to be involved in osmotic stress responses (Mahajan et al., 2006; Xiang et al., 2007; Yang et al., 2008; Tripathi et al., 2009; Zhao et al., 2009; Cuéllar et al., 2010).In this report, we examined the role of the Arabidopsis CIPK21 gene in osmotic stress response by reverse genetic analysis. The loss-of-function mutant plants became hypersensitive to salt and mannitol stress conditions, suggesting that CIPK21 is involved in the regulation of osmotic stress response in Arabidopsis. These findings are further supported by an enhanced tonoplast targeting of the cytoplasmic CIPK21 through interaction with the vacuolar Ca2+ sensors CBL2 and CBL3 under salt stress condition.  相似文献   

18.
19.
During water stress, stomatal closure occurs as water tension and levels of abscisic acid (ABA) increase in the leaf, but the interaction between these two drivers of stomatal aperture is poorly understood. We investigate the dynamics of water potential, ABA, and stomatal conductance during the imposition of water stress on two drought-tolerant conifer species with contrasting stomatal behavior. Rapid rehydration of excised shoots was used as a means of differentiating the direct influences of ABA and water potential on stomatal closure. Pinus radiata (Pinaceae) was found to exhibit ABA-driven stomatal closure during water stress, resulting in strongly isohydric regulation of water loss. By contrast, stomatal closure in Callitris rhomboidea (Cupressaceae) was initiated by elevated foliar ABA, but sustained water stress saw a marked decline in ABA levels and a shift to water potential-driven stomatal closure. The transition from ABA to water potential as the primary driver of stomatal aperture allowed C. rhomboidea to rapidly recover gas exchange after water-stressed plants were rewatered, and was associated with a strongly anisohydric regulation of water loss. These two contrasting mechanisms of stomatal regulation function in combination with xylem vulnerability to produce highly divergent strategies of water management. Species-specific ABA dynamics are proposed as a central component of drought survival and ecology.By guarding the interface between plant and atmosphere, the stomata of land plants occupy a uniquely important role that connects diverse aspects of plant biology with atmospheric processes. Capitalizing upon the potential for stomata to be used to modify plant growth and survival, or as a tool for interpreting environmental change, requires a mechanistic understanding of how these tiny valves operate. Yet, an integrated understanding of stomatal control remains elusive. Foremost in this uncertainty is an explanation for how complex signals from the environment are translated into guard cell movement. A particularly challenging feature of stomatal behavior is the fact that environmental perturbation induces both physical and chemical responses within the plant and that turgor-regulated stomata are responsive to both signals. Disentangling these distinct contributions to stomatal conductance (gs) has been made more complicated by the limited communication between molecular-scaled disciplines of mutant characterization and membrane transport biology and researchers at the larger scale of plant water relations and xylem transport. As a result, two contrasting views of stomatal control exist. Molecular biologists view stomata as osmotically regulated valves uniquely responsive to plant hormone levels and the resultant movement of ions across the guard cell membranes (Schroeder et al., 2001; Roelfsema and Hedrich, 2005). By contrast, most process-based models assume a direct influence of soil water content on stomatal aperture (Buckley, 2005; Damour et al., 2010).The phytohormone abscisic acid (ABA) is seen as a cornerstone of stomatal function because it has been shown to trigger responses in guard cell membrane channels and transporters that cause a reduction in guard cell turgor, thereby closing stomata. ABA-mediated stomatal closure in seed plants (but not in ferns and lycophytes; Brodribb and McAdam, 2011) is broadly accepted as the explanation for stomatal closure during water stress (Zhang and Davies, 1989; Bauer et al., 2013); yet, there are very few studies that show a good correlation between the level of ABA and gs during water stress in the field. The traditional explanation for this lack of a strong relationship suggests that ABA is a root-derived hormone that is delivered to the leaf in the transpiration stream (Zhang et al., 1987; Davies and Zhang, 1991) and hence that the xylem ABA flux, rather than the leaf level of ABA, should dictate the intensity of the stomatal response to soil drying (Tardieu et al., 1992; Tardieu and Davies, 1993). The flux-based model for ABA action in the leaf remains the most widely used interpretation of how stomata sense and respond to drying soil, despite the fact that there is mounting evidence for significant ABA synthesis in the leaf and guard cells, and short term responses to ABA that cannot be explained by xylem transport (Christmann et al., 2005; Lee et al., 2006; Georgopoulou and Milborrow, 2012). Furthermore, the ABA flux approach has never been successfully applied to explain variation in transpiration in trees (Sperry, 2000; Cochard et al., 2002), suggesting that there may be some benefit in reexamining some of the principles and assumptions used to link water stress, ABA, and transpiration.Here, we examine the dynamics of stomatal closure, leaf ABA levels, and xylem tension during the gradual imposition of water stress upon two conifer species, Pinus radiata and Callitris rhomboidea, known for having contrasting stomatal responses to desiccation. Our primary aim is to separate the interacting effects of ABA and water tension on guard cell turgor pressure and stomatal diffusive conductance and hence to reveal the relative importance of water tension and ABA levels during drought as effectors of stomatal closure. Conifers are particularly suitable for identifying different closing signals because they do not appear to produce hydropassive stomatal movements (McAdam and Brodribb, 2012). This makes them ideal for examining the direct effects of ABA and water tension without the mechanical interactions between subsidiary cells and guard cells (Franks and Farquhar, 2007) that greatly complicate the mechanics of angiosperm stomatal movements. Both conifer species examined grow naturally in low rainfall habitats, but P. radiata is strongly isohydric (meaning that stomata close in a very narrow range of leaf hydration), while C. rhomboidea is anisohydric (meaning that stomata have a relatively low sensitivity to leaf hydration).  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号