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 175 毫秒
1.
2.
3.
Motivated by studies suggesting that the stomata of ferns and lycophytes do not conform to the standard active abscisic acid (ABA) -mediated stomatal control model, we examined stomatal behavior in a conifer species (Metasequoia glyptostroboides) that is phylogenetically midway between the fern and angiosperm clades. Similar to ferns, daytime stomatal closure in response to moderate water stress seemed to be a passive hydraulic process in M. glyptostroboides immediately alleviated by rehydrating excised shoots. Only after prolonged exposure to more extreme water stress did active ABA-mediated stomatal closure become important, because foliar ABA production was triggered after leaf turgor loss. The influence of foliar ABA on stomatal conductance and stomatal aperture was highly predictable and additive with the passive hydraulic influence. M. glyptostroboides thus occupies a stomatal behavior type intermediate between the passively controlled ferns and the characteristic ABA-dependent stomatal closure described in angiosperm herbs. These results highlight the importance of considering phylogeny as a major determinant of stomatal behavior.Stomata regulate parallel diffusive paths of water and carbon dioxide between leaves and the atmosphere, thus assuming a governing role over the processes of transpiration and photosynthesis. Guard cell movements that open and close stomata are increasingly characterized (similarly to animal cells) as being mediated by rapid changes in the polarization state of membranes (Schroeder et al., 2001; Hedrich, 2012; Hills et al., 2012). Despite this membrane-dominated view of stomatal function, the critical goal of modeling stomatal behavior to render predictions of transpiration and photosynthesis typically relies on a hydraulic framework built around the direct impact of leaf hydration on epidermal and guard cell turgor (Buckley, 2005; Damour et al., 2010). Although recent advances in modeling the ionic balance of guard cells (Chen et al., 2012; Hills et al., 2012) yield predictions of stomatal aperture, no macro-scale stomatal model has been able to predict stomatal conductance from the perspective of ion movements into and out of guard cells; however, the effects of key components, such as light, carbon dioxide, and abscisic acid (ABA), on membrane polarization have been studied in detail. Reconciling the dynamics of leaf-scale and canopy-scale transpiration with physical and chemical processes at the guard cell remains a major challenge.Among the obstacles preventing the formulation of a large-scale transpiration model based on membrane ion transport is the fact that much of the characterization of guard cell membrane processes has been confined to a handful of small, ruderal, herbaceous angiosperms. Although species like Arabidopsis (Arabidopsis thaliana) provide the ideal molecular system for identifying guard cell signal transduction pathways, most of these model species are of little agricultural relevance and being herbaceous, poor physical analogs for the tree species that dominate terrestrial gas exchange. Hence, there is a need to understand whether the same principles governing stomatal control in angiosperm herbs, like Arabidopsis, equally apply to plants that dominate forests and agricultural production. Recent studies suggest that there are important differences in the ion-transport machinery among vascular plants, and despite the presence of potential guard cell signaling pathways throughout the plant kingdom (Dreyer et al., 2012; Brodribb and McAdam, 2013b; Chater et al., 2013), there is evidence of a systematic shift in the behavior of stomata among vascular plants (Doi et al., 2006; Brodribb and McAdam, 2011; McAdam and Brodribb, 2012a). In particular, the critical closing tendency of stomata during leaf water deficit seems to have evolved from a passive process mediated directly by water potential (passive hydraulic) to an active process controlled by the extrusion of anions from guard cells (active closure; Brodribb and McAdam, 2011). The stomata of ferns and lycophytes predictably respond to plant water deficit as passive hydraulic valves, closing rapidly on dehydration and opening on rehydration (Brodribb and McAdam, 2011; McAdam and Brodribb, 2012a). Despite the stomata in these lineages only ever showing functionally passive responses to changes in leaf water status (Brodribb and McAdam, 2011; McAdam and Brodribb, 2012a, 2013), some have challenged the concept of a passive origin of stomatal control in vascular plants by showing a conserved activity of key genes involved in active stomatal responses (Ruszala et al., 2011). In seed plants, the closure of stomata in response to water deficit is mediated by augmented levels of ABA, which leads to a depolarization of guard cell membranes triggering osmotic ion efflux and a loss of guard cell turgor (Mittelheuser and Van Steveninck, 1969; Thiel et al., 1992; Geiger et al., 2009, 2011; Bauer et al., 2013). In light of this variation in stomatal control, it seems that a key step to finding a general model for stomatal behavior would be to understand the interactions between active and passive processes in the stomatal movements of major lineages of plants.Conifers contribute significantly to global transpiration and productivity and also seem to have a stomatal control system that is somewhat different from model angiosperm herbs. These distinctions include insensitivity to elevated carbon dioxide (Beadle et al., 1979; Morison and Jarvis, 1983; Brodribb et al., 2009); a lack of epidermal mechanical advantage, resulting in no Ivanov effect (the increase in transpiration from a leaf after excision or exposure to low humidity; Huber, 1923; Stålfelt, 1944; McAdam and Brodribb, 2012a), likely because of heavily lignified dorsal walls (the walls closest to the epidermal cells; Sack, 1987), and a very high length-to-width ratio of open stomatal pores (Copeland, 1902). Furthermore, recent research suggests that different conifer species depend more or less on ABA as an agent of stomatal closure during extended periods of water stress (Brodribb and McAdam, 2013a). The apparent lack of epidermal mechanical advantage in conifer stomata provides an unusual opportunity to examine the impacts of changing leaf water content and evaporation on stomatal conductance and guard cell turgor without the confusing Ivanov effect produced by changes in the ratio of epidermal and guard cell turgor pressure (Raschke, 1970). Manipulating the hydration status of the leaf, thus, allows quantification of the interacting influences of leaf water potential (Ψl) and ion transport on stomatal aperture (Brodribb and McAdam, 2013a).Our aim in this study was to determine under what conditions passive (hydraulic) and active (ABA mediated) closures of stomata were important in a representative conifer species. We assumed that, in the absence of a mechanical interaction from the epidermis, it would be possible to characterize both dynamic and steady-state stomatal behavior based on intrinsic leaf properties of ABA sensitivity, hydraulic conductance, capacitance, and hydraulic vulnerability. We chose the conifer Metasequoia glyptostroboides (Cupressaceae) as our subject, because it has leaf characteristics within the range of deciduous angiosperm trees; also, it is one of the few conifer species where stomata are sufficiently visible (unoccluded by waxes) to observe stomatal responses in the isolated epidermis.  相似文献   

4.
5.
6.
Developmental responses associated with end-of-day far-red light (EOD-FR) signaling were investigated in maize (Zea mays subspecies mays) seedlings. A survey of genetically diverse inbreds of temperate and tropical/semitropical origins, together with teosinte (Zea mays subspecies parviglumis) and a modern hybrid, revealed distinct elongation responses. A mesocotyl elongation response to the EOD-FR treatment was largely absent in the tropical/semitropical lines, but both hybrid and temperate inbred responses were of the same magnitude as in teosinte, suggesting that EOD-FR-mediated mesocotyl responses were not lost during the domestication or breeding process. The genetic architecture underlying seedling responses to EOD-FR was investigated using the intermated B73 × Mo17 mapping population. Among the different quantitative trait loci identified, two were consistently detected for elongation and responsiveness under EOD-FR, but none were associated with known light signaling loci. The central role of phytochromes in mediating EOD-FR responses was shown using a phytochromeB1 phytochromeB2 (phyB1 phyB2) mutant series. Unlike the coleoptile and first leaf sheath, EOD-FR-mediated elongation of the mesocotyl appears predominantly controlled by gibberellin. EOD-FR also reduced abscisic acid (ABA) levels in the mesocotyl for both the wild type and phyB1 phyB2 double mutants, suggesting a FR-mediated but PHYB-independent control of ABA accumulation. EOD-FR elongation responses were attenuated in both the wild type and phyB1 phyB2 double mutants when a chilling stress was applied during the dark period, concomitant with an increase in ABA levels. We present a model for the EOD-FR response that integrates light and hormonal control of seedling elongation.Plants utilize a complex network of photoreceptors to monitor the spectral quality, fluence, direction, and duration of light (Smith, 1995). These photosensory pigments include phytochromes that sense red light (R; 580–690 nm) and far-red light (FR; 690–800 nm) and the cryptochromes, phototropins, and zeitlupes for blue light (380–495 nm) and UV-A light (320–380 nm). The light reflected and transmitted by the vegetation creates a canopy characterized by reductions in both the R-to-FR ratio (R:FR) and the photosynthetically active radiation (400–700 nm). This light environment induces adaptive biochemical and morphological responses known as the shade avoidance syndrome (Smith and Whitelam, 1997). These responses can be induced early in development, before canopy closure, through FR reflected from adjacent neighbor plants (Ballare et al., 1990) or from low-lying weeds (Rajcan and Swanton, 2001), which can negatively impact yields in maize (Zea mays subspecies mays; Rajcan et al., 2004), even if only present early in the growing season (Liu et al., 2009).R:FR signals are transduced by the phytochrome family of photoreceptors (Franklin and Whitelam, 2007b). In rice (Oryza sativa) and sorghum (Sorghum bicolor), three genes constitute the phytochrome family: PhytochromeA (PhyA), PhytochromeB (PhyB), and PhytochromeC (PhyC). In maize, an ancient alloploidization has doubled the family size to six: PhyA1, PhyA2, PhyB1, PhyB2, PhyC1, and PhyC2 (Sheehan et al., 2004). Although many similarities are apparent between maize and Arabidopsis (Arabidopsis thaliana) light response, there are significant differences between members of the phytochrome gene family in copy number and selection pressures that have resulted in the divergence of phytochrome signaling networks (Sawers et al., 2005; Sheehan et al., 2007). Thus far, only three phytochrome mutants have been characterized in maize: elongated mesocotyl1 (elm1), phyB1, and phyB2. The elm1 mutant carries a mutation in phytochromobilin synthase, necessary for the biosynthesis of the chromophore common to all phytochromes (Sawers et al., 2004). The mutation severely reduces the total phytochrome pool, but the weak nature of the allele enables a partial responsiveness to R and FR (Markelz et al., 2003). At maturity, elm1 mutants have elongated internodes, pale green leaves, and flower early (Sawers et al., 2002). Mutations at phyB1 and phyB2 also impair light signal transduction. At maturity, both PHYB1 and PHYB2 contribute to plant height, stem diameter, and sheath-internode length, but PHYB2 predominates in the control of flowering (Sheehan et al., 2007). Like the sorghum and rice phyB mutants (Childs et al., 1997; Takano et al., 2005; Kebrom et al., 2010), both elm1 and phyB1 phyB2 double mutants constitutively display several traits associated with low R:FR response (Sawers et al., 2002; Markelz et al., 2003; Sheehan et al., 2007).In Arabidopsis, R/FR-mediated responses are developmentally complex and involve the PIF proteins (Duek and Fankhauser, 2005) and many hormones including auxins (Tao et al., 2008), ethylene (Khanna et al., 2007), jasmonate (Moreno et al., 2009), and GA (Djakovic-Petrovic et al., 2007). In particular, there is a direct interaction between PIF and DELLA proteins that govern phytochrome-mediated elongation (de Lucas et al., 2008; Feng et al., 2008; Lorrain et al., 2008). DELLA proteins also regulate FR inhibition of germination by abscisic acid (ABA; Piskurewicz et al., 2009), suggesting an interaction between the PIFs and ABA signaling. Complex cross talk between light and temperature has also been reported (Franklin, 2009). In Arabidopsis, colder temperatures can repress the early-flowering phenotype of the phyB mutant (Halliday et al., 2003). These studies suggest a complex interplay between light, hormone, and temperature to fine-tune the elongation response.The end-of-day far-red light (EOD-FR) treatment consists of a pulse of FR given at subjective dusk (Kasperbauer, 1971) and triggers a circadian clock-gated response (Salter et al., 2003). EOD-FR treatments result in a minimal pool of active Pfr during the dark period (Fankhauser and Casal, 2004), and plants submitted to daily treatments display similar developmental responses to those elicited by a continuous photoperiod with low R:FR (Smith, 1994). One of the key features that contributed to the discovery of the phytochromes is the photoreversibility of the response (Borthwick et al., 1952). These low-fluence responses (LFRs) are induced or repressed by alternating R and FR treatments (Mancinelli, 1994). The LFR nature of EOD-FR in maize was previously demonstrated in 5-d-old mesocotyl and coleoptile tissues (Gorton and Briggs, 1980). The EOD-FR treatment offers several advantages over growing plants in continuous low R:FR, including exposing plants to relatively brief treatment periods, thus potentially reducing genotype × environment effects. It also facilitates kinetic assays of phytochrome response, as treatments are limited to a single point in the diurnal cycle and can be delivered at any stage in plant development. Finally, as relatively low fluences of light are needed to saturate EOD-FR responses, large populations of seedlings can be screened without the need for large numbers of FR light-emitting diodes (LEDs) or sophisticated light chambers.Here, we have examined EOD-FR-mediated responses in maize and its closest relative, teosinte (Zea mays subspecies parviglumis). A survey of genetically diverse maize and teosinte accessions revealed extensive tissue-specific variations in mesocotyl, coleoptile, and first leaf sheath elongation. EOD-FR responses were greatly attenuated in tropical/semitropical (TS) accessions but present in teosinte, temperate inbreds, and a modern commercial hybrid, suggesting that the EOD-FR response is plastic in Z. mays. To investigate the genetic regulation underlying seedling responses to EOD-FR, we performed a quantitative trait locus (QTL) analysis using the intermated B73 × Mo17 (IBM) recombinant inbred population. We identified several QTLs that regulate mesocotyl and first leaf sheath response to EOD-FR and show that these QTLs mediate tissue-specific responses. The phyB1 phyB2 mutant series was also evaluated, indicating that the two PhyB paralogs are largely redundant in mediating the EOD-FR response. Pharmacological assays revealed a major role for GA in promoting mesocotyl, but not coleoptile or first leaf sheath, elongation in response to EOD-FR treatments. In contrast, EOD-FR reduced mesocotyl ABA levels. A chill treatment (10°C) applied during dark breaks attenuated EOD-FR elongation responses. Based on these observations, we discuss a model that integrates temperature, light, and hormonal inputs in the regulation of mesocotyl elongation.  相似文献   

7.
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   

8.
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   

9.
The rate of gas exchange in plants is regulated mainly by stomatal size and density. Generally, higher densities of smaller stomata are advantageous for gas exchange; however, it is unclear what the effect of an extraordinary change in stomatal size might have on a plant’s gas-exchange capacity. We investigated the stomatal responses to CO2 concentration changes among 374 Arabidopsis (Arabidopsis thaliana) ecotypes and discovered that Mechtshausen (Me-0), a natural tetraploid ecotype, has significantly larger stomata and can achieve a high stomatal conductance. We surmised that the cause of the increased stomatal conductance is tetraploidization; however, the stomatal conductance of another tetraploid accession, tetraploid Columbia (Col), was not as high as that in Me-0. One difference between these two accessions was the size of their stomatal apertures. Analyses of abscisic acid sensitivity, ion balance, and gene expression profiles suggested that physiological or genetic factors restrict the stomatal opening in tetraploid Col but not in Me-0. Our results show that Me-0 overcomes the handicap of stomatal opening that is typical for tetraploids and achieves higher stomatal conductance compared with the closely related tetraploid Col on account of larger stomatal apertures. This study provides evidence for whether larger stomatal size in tetraploids of higher plants can improve stomatal conductance.Gas exchange is a vital activity for higher plants that take up atmospheric CO2 and release oxygen and water vapor through epidermal stomatal pores. Gas exchange affects CO2 uptake, photosynthesis, and biomass production (Horie et al., 2006; Evans et al., 2009; Tanaka et al., 2014). Stomatal conductance (gs) is used as an indicator of gas-exchange capacity (Franks and Farquhar, 2007). Maximum stomatal conductance (gsmax) is controlled mainly by stomatal size and density, two parameters that change with environmental conditions and are negatively correlated with each other (Franks et al., 2009).Given a constant total stomatal pore area, large stomata are generally disadvantageous for gas exchange compared with smaller stomata, because the greater pore depth in larger stomata increases the distance that gas molecules diffuse through. This increased distance is inversely proportional to gsmax (Franks and Beerling, 2009). The fossil record indicates that ancient plants had small numbers of large stomata when atmospheric CO2 levels were high, and falling atmospheric [CO2] induced a decrease in stomatal size and an increase in stomatal density to increase gs for maximum carbon gain (Franks and Beerling, 2009). The positive relationship between a high gs and numerous small stomata also holds true among plants living today under various environmental conditions (Woodward et al., 2002; Galmés et al., 2007; Franks et al., 2009). Additionally, the large stomata of several plant species (e.g. Vicia faba and Arabidopsis [Arabidopsis thaliana]) are often not effective for achieving rapid changes in gs, due to slower solute transport to drive movement caused by their lower membrane surface area-to-volume ratios (Lawson and Blatt, 2014).Stomatal size is strongly and positively correlated with genome size (Beaulieu et al., 2008; Franks et al., 2012; Lomax et al., 2014). Notably, polyploidization causes dramatic increases in nucleus size and stomatal size (Masterson, 1994; Kondorosi et al., 2000). In addition to the negative effects of large stomata on gas exchange (Franks et al., 2009), polyploids may have another disadvantage; del Pozo and Ramirez-Parra (2014) showed that artificially induced tetraploids of Arabidopsis have a reduced stomatal density (stomatal number per unit of leaf area) and a lower stomatal index (stomatal number per epidermal cell number). Moreover, tetraploids of Rangpur lime (Citrus limonia) and Arabidopsis have lower transpiration rates and changes in the expression of genes involved in abscisic acid (ABA), a phytohormone that induces stomatal closure (Allario et al., 2011; del Pozo and Ramirez-Parra, 2014). On the other hand, an increase in the ploidy level of Festuca arundinacea results in an increase in the CO2-exchange rate (Byrne et al., 1981); hence, polyploids may not necessarily have a reduced gas-exchange capacity.Natural accessions provide a wide range of information about mechanisms for adaptation, regulation, and responses to various environmental conditions (Bouchabke et al., 2008; Brosché et al., 2010). Arabidopsis, which is distributed widely throughout the Northern Hemisphere, has great natural variation in stomatal anatomy (Woodward et al., 2002; Delgado et al., 2011). Recently, we investigated leaf temperature changes in response to [CO2] in a large number of Arabidopsis ecotypes (374 ecotypes; Takahashi et al., 2015) and identified the Mechtshausen (Me-0) ecotype among ecotypes with low CO2 responsiveness; Me-0 had a comparatively low leaf temperature, implying a high transpiration rate. In this study, we revealed that Me-0 had a higher gs than the standard ecotype Columbia (Col), despite having tetraploid-dependent larger stomata. Notably, the gs of Me-0 was also higher than that of tetraploid Col, which has stomata as large as those of Me-0. This finding resulted from Me-0 having a higher gs-to-gsmax ratio due to more opened stomata than tetraploid Col. In addition, there were differences in ABA responsiveness, ion homeostasis, and gene expression profiles in guard cells between Me-0 and tetraploid Col, which may influence their stomatal opening. Despite the common trend of smaller stomata with higher gas-exchange capacity, the results with Me-0 confirm the theoretical possibility that larger stomata can also achieve higher stomatal conductance if pore area increases sufficiently.  相似文献   

10.
11.
12.
13.
14.
15.
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   

16.
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   

17.
18.
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   

19.
20.
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   

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