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.
The photosynthetic assimilation of CO2 in C4 plants is potentially limited by the enzymatic rates of Rubisco, phosphoenolpyruvate carboxylase (PEPc), and carbonic anhydrase (CA). Therefore, the activity and kinetic properties of these enzymes are needed to accurately parameterize C4 biochemical models of leaf CO2 exchange in response to changes in CO2 availability and temperature. There are currently no published temperature responses of both Rubisco carboxylation and oxygenation kinetics from a C4 plant, nor are there known measurements of the temperature dependency of the PEPc Michaelis-Menten constant for its substrate HCO3, and there is little information on the temperature response of plant CA activity. Here, we used membrane inlet mass spectrometry to measure the temperature responses of Rubisco carboxylation and oxygenation kinetics, PEPc carboxylation kinetics, and the activity and first-order rate constant for the CA hydration reaction from 10°C to 40°C using crude leaf extracts from the C4 plant Setaria viridis. The temperature dependencies of Rubisco, PEPc, and CA kinetic parameters are provided. These findings describe a new method for the investigation of PEPc kinetics, suggest an HCO3 limitation imposed by CA, and show similarities between the Rubisco temperature responses of previously measured C3 species and the C4 plant S. viridis.Biochemical models of photosynthesis are often used to predict the effect of environmental conditions on net rates of leaf CO2 assimilation (Farquhar et al., 1980; von Caemmerer, 2000, 2013; Walker et al., 2013). With climate change, there is increased interest in modeling and understanding the effects of changes in temperature and CO2 concentration on photosynthesis. The biochemical models of photosynthesis are primarily driven by the kinetic properties of the enzyme Rubisco, the primary carboxylating enzyme of the C3 photosynthetic pathway, catalyzing the reaction of ribulose-1,5-bisphosphate (RuBP) with either CO2 or oxygen. However, the CO2-concentrating mechanism in C4 photosynthesis utilizes carbonic anhydrase (CA) to help maintain the chemical equilibrium of CO2 with HCO3 and phosphoenolpyruvate carboxylase (PEPc) to catalyze the carboxylation of phosphoenolpyruvate (PEP) with HCO3. These reactions ultimately provide the elevated levels of CO2 to the compartmentalized Rubisco (Edwards and Walker, 1983). In C4 plants, it has been demonstrated that PEPc, Rubisco, and CA can limit rates of CO2 assimilation and influence the efficiency of the CO2-concentrating mechanism (von Caemmerer, 2000; von Caemmerer et al., 2004; Studer et al., 2014). Therefore, accurate modeling of leaf photosynthesis in C4 plants in response to future climatic conditions will require temperature parameterizations of Rubisco, PEPc, and CA kinetics from C4 species.Modeling C4 photosynthesis relies on the parameterization of both PEPc and Rubisco kinetics, making it more complex than for C3 photosynthesis (Berry and Farquhar, 1978; von Caemmerer, 2000). However, the activity of CA is not included in these models, as it is assumed to be nonlimiting under most conditions (Berry and Farquhar, 1978; von Caemmerer, 2000). This assumption is implemented by modeling PEPc kinetics as a function of CO2 partial pressure (pCO2) and not HCO3 concentration, assuming CO2 and HCO3 are in chemical equilibrium. However, there are questions regarding the amount of CA activity needed to sustain rates of C4 photosynthesis and if CO2 and HCO3 are in equilibrium (von Caemmerer et al., 2004; Studer et al., 2014).The most common steady-state biochemical models of photosynthesis are derived from the Michaelis-Menten models of enzyme activity (von Caemmerer, 2000), which are driven by the Vmax and the Km. Both of these parameters need to be further described by their temperature responses to be used to model photosynthesis in response to temperature. However, the temperature response of plant CA activity has not been completed above 17°C, and there is no known measured temperature response of Km HCO3 for PEPc (KP). Alternatively, Rubisco has been well studied, and there are consistent differences in kinetic values between C3 and C4 species at 25°C (von Caemmerer and Quick, 2000; Kubien et al., 2008), but the temperature responses, including both carboxylation and oxygenation reactions, have only been performed in C3 species (Badger and Collatz, 1977; Jordan and Ogren, 1984; Bernacchi et al., 2001, 2002; Walker et al., 2013).Here, we present the temperature dependency of Rubisco carboxylation and oxygenation reactions, PEPc kinetics for HCO3, and CA hydration from 10°C to 40°C from the C4 species Setaria viridis (succession no., A-010) measured using membrane inlet mass spectrometry. Generally, the 25°C values of the Rubisco parameters were similar to previous measurements of C4 species. The temperature response of the maximum rate of Rubisco carboxylation (Vcmax) was high compared with most previous measurements from both C3 and C4 species, and the temperature response of the Km for oxygenation (KO) was low compared with most previously measured species. Taken together, the modeled temperature responses of Rubisco activity in S. viridis were similar to the previously reported temperature responses of some C3 species. Additionally, the temperature response of the maximum rate of PEPc carboxylation (Vpmax) was similar to previous measurements. However, the temperature response of KP was lower than what has been predicted (Chen et al., 1994). For CA, deactivation of the hydration activity was observed above 25°C. Additionally, models of CA and PEPc show that CA activity limits HCO3 availability to PEPc above 15°C, suggesting that CA limits PEP carboxylation rates in S. viridis when compared with the assumption that CO2 and HCO3 are in full chemical equilibrium.  相似文献   

2.
The mitochondrial electron transport chain includes an alternative oxidase (AOX) that is hypothesized to aid photosynthetic metabolism, perhaps by acting as an additional electron sink for photogenerated reductant or by dampening the generation of reactive oxygen species. Gas exchange, chlorophyll fluorescence, photosystem I (PSI) absorbance, and biochemical and protein analyses were used to compare respiration and photosynthesis of Nicotiana tabacum ‘Petit Havana SR1’ wild-type plants with that of transgenic AOX knockdown (RNA interference) and overexpression lines, under both well-watered and moderate drought-stressed conditions. During drought, AOX knockdown lines displayed a lower rate of respiration in the light than the wild type, as confirmed by two independent methods. Furthermore, CO2 and light response curves indicated a nonstomatal limitation of photosynthesis in the knockdowns during drought, relative to the wild type. Also relative to the wild type, the knockdowns under drought maintained PSI and PSII in a more reduced redox state, showed greater regulated nonphotochemical energy quenching by PSII, and displayed a higher relative rate of cyclic electron transport around PSI. The origin of these differences may lie in the chloroplast ATP synthase amount, which declined dramatically in the knockdowns in response to drought. None of these effects were seen in plants overexpressing AOX. The results show that AOX is necessary to maintain mitochondrial respiration during moderate drought. In its absence, respiration rate slows and the lack of this electron sink feeds back on the photosynthetic apparatus, resulting in a loss of chloroplast ATP synthase that then limits photosynthetic capacity.The plant mitochondrial electron transport chain (ETC) is bifurcated such that electrons in the ubiquinone pool partition between the cytochrome (cyt) pathway (consisting of Complex III, cyt c, and Complex IV) and alternative oxidase (AOX; Finnegan et al., 2004; Millar et al., 2011; Vanlerberghe, 2013). AOX directly couples ubiquinol oxidation with O2 reduction to water. This reduces the energy yield of respiration because, unlike Complexes III and IV, AOX is not proton pumping. Hence, AOX is an electron sink, the capacity of which is little encumbered by rates of ATP turnover. In this way, AOX might be well suited to prevent cellular over-reduction. Supporting this, transgenic Nicotiana tabacum leaves with suppressed amounts of AOX have increased concentrations of mitochondrial-localized superoxide radical (O2) and nitric oxide, the products that can arise when an over-reduced ETC results in electron leakage to O2 or nitrite (Cvetkovska and Vanlerberghe, 2012, 2013).In angiosperms, AOX is encoded by a small gene family (Considine et al., 2002). In Arabidopsis (Arabidopsis thaliana), mutation or knockdown of the stress-responsive AOX1a gene family member dramatically reduces AOX protein and the capacity of the AOX respiration pathway to consume O2. Several studies have shown that this loss of AOX capacity in Arabidopsis aox1a plants affected processes such as growth, carbon and energy metabolism, and/or the cellular network of reactive oxygen species (ROS) scavengers (Fiorani et al., 2005; Umbach et al., 2005; Watanabe et al., 2008; Giraud et al., 2008; Skirycz et al., 2010). However, in studies in which respiration was measured, it was consistently reported that the lack of AOX capacity had no significant impact on the respiration rate in the dark (RD; Umbach et al., 2005; Giraud et al., 2008; Strodtkötter et al., 2009; Florez-Sarasa et al., 2011; Yoshida et al., 2011b; Gandin et al., 2012). The exceptions are two reports that RD was actually higher in aox1a than in the wild type under some conditions (Watanabe et al., 2008; Vishwakarma et al., 2014). To our knowledge, how the lack of AOX affects respiration rate in the light (RL) is not reported in Arabidopsis or other species.Numerous studies have established the importance of mitochondrial metabolism in the light to optimize photosynthesis (Hoefnagel et al., 1998; Raghavendra and Padmasree, 2003). In recent years, the potential importance of specifically AOX respiration during photosynthesis has been examined using the Arabidopsis aox1a plants (Giraud et al., 2008; Strodtkötter et al., 2009; Zhang et al., 2010; Florez-Sarasa et al., 2011; Yoshida et al., 2011a, 2011b). In general, these studies reported small perturbations of photosynthesis in standard-grown aox1a plants, including slightly lower rates of CO2 uptake or O2 release (Gandin et al., 2012; Vishwakarma et al., 2014), slightly higher rates of cyclic electron transport (CET; Yoshida et al., 2011b), and slightly increased susceptibility to photoinhibition after a high light treatment (Florez-Sarasa et al., 2011). Generally, these studies concluded that aox1a plants exhibit a biochemical limitation of photosynthesis, in line with the hypothesis that AOX serves as a sink for excess photogenerated reducing power, with the reductant likely reaching the mitochondrion via the malate valve (Noguchi and Yoshida, 2008; Taniguchi and Miyake, 2012). Similar to these Arabidopsis studies, we recently reported that well-watered N. tabacum AOX knockdowns grown at moderate irradiance display a slight reduced rate of photosynthesis (approximately 10%–15%) when measured at high irradiance. However, we established that the lower photosynthetic rate was the result of a stomatal rather than biochemical limitation of photosynthesis, and provided evidence that this stomatal limitation resulted from disrupted nitric oxide homeostasis within the guard cells of AOX knockdown plants (Cvetkovska et al., 2014).Drought is a common abiotic stress that can substantially curtail photosynthesis because stomatal closure, meant to conserve water, also restricts CO2 availability to the Calvin cycle. Besides this well established stomatal limitation of photosynthesis, there may also be water deficit-sensitive biochemical components that contribute to the reduction of photosynthesis during drought. However, the nature of this biochemical limitation and the degree to which it contributes to the curtailment of photosynthesis during drought remain areas of active debate (Flexas et al., 2004; Lawlor and Tezara, 2009; Pinheiro and Chaves, 2011). Additional factors, such as patchy stomatal closure (Sharkey and Seemann, 1989; Gunasekera and Berkowitz, 1992) or changes in the conductance to CO2 of mesophyll cells (Perez-Martin et al., 2009), can further complicate analyses of photosynthesis during drought.Metabolism can experience energy imbalances, when there is a mismatch between rates of synthesis and rates of utilization of ATP and/or NADPH, and the importance of mechanisms to minimize such imbalances has been emphasized (Cruz et al., 2005; Kramer and Evans, 2011; Vanlerberghe, 2013). For example, such imbalances may occur in the chloroplast when the use of ATP and NADPH by the Calvin cycle does not keep pace with the harvesting of light energy (Hüner et al., 2012). This can result in excess excitation energy that can damage photosynthetic components, perhaps through the generation of ROS (Asada, 2006; Noctor et al., 2014). Such a scenario has been hypothesized to underlie the development of the biochemical limitations of photosynthesis reported during drought (Lawlor and Tezara, 2009).In this study, we find that N. tabacum AOX knockdowns show a compromised rate of mitochondrial respiration in the light during moderate drought. This corresponds with a strong nonstomatal limitation of photosynthesis in these plants relative to the wild type, and we describe a biochemical basis for this photosynthetic limitation. The results indicate that AOX is a necessary electron sink to support photosynthesis during drought, a condition when the major photosynthetic electron sink, the Calvin cycle, is becoming limited by CO2 availability.  相似文献   

3.
Bypassing the photorespiratory pathway is regarded as a way to increase carbon assimilation and, correspondingly, biomass production in C3 crops. Here, the benefits of three published photorespiratory bypass strategies are systemically explored using a systems-modeling approach. Our analysis shows that full decarboxylation of glycolate during photorespiration would decrease photosynthesis, because a large amount of the released CO2 escapes back to the atmosphere. Furthermore, we show that photosynthesis can be enhanced by lowering the energy demands of photorespiration and by relocating photorespiratory CO2 release into the chloroplasts. The conductance of the chloroplast membranes to CO2 is a key feature determining the benefit of the relocation of photorespiratory CO2 release. Although our results indicate that the benefit of photorespiratory bypasses can be improved by increasing sedoheptulose bisphosphatase activity and/or increasing the flux through the bypass, the effectiveness of such approaches depends on the complex regulation between photorespiration and other metabolic pathways.In C3 plants, the first step of photosynthesis is the fixation of CO2 by ribulose bisphosphate (RuBP). For every molecule of CO2 fixed, this reaction produces two molecules of a three-carbon acid, i.e., 3-phosphoglycerate (PGA), and is catalyzed by the Rubisco enzyme. A small portion of the carbon in PGA is used for the production of Suc and starch, whereas the remainder (i.e. five-sixths) is used for the regeneration of RuBP (Fig. 1). The regeneration of the Rubisco substrate RuBP in the Calvin-Benson-Bassham (CBB) cycle ensures that ample RuBP is available for carbon fixation (Bassham, 1964; Wood, 1966; Beck and Hopf, 1982). Rubisco is a bifunctional enzyme that catalyzes not only RuBP carboxylation but also RuBP oxygenation (Spreitzer and Salvucci, 2002). RuBP oxygenation generates only one molecule of PGA and one molecule of 2-phosphoglycolate (P-Gly; Ogren, 1984). The photorespiratory pathway converts this P-Gly back to RuBP in order to maintain the CBB cycle.Open in a separate windowFigure 1.Schematic representation of the C3 photosynthesis kinetic model with three different photorespiratory bypass pathways. The bypass described by Kebeish et al. (2007) is indicated in blue, the bypass described by Maier et al. (2012) in pink, and the bypass described by Carvalho et al. (2011) in green. The original photorespiratory pathway is marked in orange, and CO2 released from photorespiration (including the original pathway and bypass pathways) is indicated in red. 2PGA, 2-Phosphoglyceric acid; ASP, Asp; CIT, citrate; ICIT, isocitrate; PGA, 3-phosphoglycerate; DPGA, glycerate-1,3-bisphosphate; GAP, glyceraldehyde 3-phosphate; DHAP, dihydroxyacetone phosphate; SBP, sedoheptulose-1,7-bisphosphate; S7P, sedoheptulose-7-phosphate; Ri5P, ribose-5-phosphate; Ru5P, ribulose-5-phosphate; FBP, fructose-1,6-bisphosphatase; F6P, fructose 6-phosphate; Xu5P, xylulose-5-phosphate; G6P, glucose-6-phosphate; G1P, glucose-1-phosphate; ADPG, ADP-glucose; F26BP, fructose-2,6-bisphosphate; UDPG, uridine diphosphate glucose; SUCP, sucrose-6F-phosphate; SUC, Suc; PEP, phosphoenolpyruvate; OAA, oxaloacetate; PGCA, phosphoglycolate; GCA, glycolate; GOA, glyoxylate; GCEA, glycerate; MAL, malate; PYR, pyruvate; GLU, glutamate; KG, alfa-ketoglutarate; GLN, Gln; HPR, hydroxypyruvate; RuBP, ribulose bisphosphate; SER, Ser; GLY, Gly; TS, tartronic semialdehyde.In higher plants, P-Gly is dephosphorylated to glycolate, which is transferred into the peroxisomes, where it is oxidized to hydrogen peroxide and glyoxylate. Then, glyoxylate is aminated to produce Gly, which is subsequently transferred to the mitochondria. There, two molecules of Gly are converted into one Ser plus one CO2 and one NH3 (Ogren, 1984; Peterhansel et al., 2010). The Ser is ultimately converted back to PGA (Tolbert, 1997). CO2 and NH3 are gasses that can escape to the atmosphere (Sharkey, 1988; Kumagai et al., 2011), and the loss of carbon and nitrogen essential for biomass accumulation will decrease the efficiency of photosynthesis and plant growth (Zhu et al., 2010). Fortunately, both substances are partially reassimilated in the chloroplast, but this results in decreased photosynthetic energy efficiency. At 25°C and current atmospheric CO2 concentrations, approximately 30% of the carbon fixed in C3 photosynthesis may be lost via photorespiration and the size of this loss increases with temperature (Sharkey, 1988; Zhu et al., 2010). As a result, photorespiration has been regarded as a pathway that could be altered to improve photosynthetic efficiency (Zelitch and Day, 1973; Oliver, 1978; Ogren, 1984; Zhu et al., 2008, 2010).There are several approaches that may be used to alter photorespiration to improve photosynthetic efficiency. First, it might be possible to increase the specificity of Rubisco to CO2 versus oxygen (Sc/o; Dhingra et al., 2004; Spreitzer et al., 2005; Whitney and Sharwood, 2007). However, previous studies have shown that there is an inverse correlation between Sc/o and the maximum carboxylation rate of Rubisco (Jordan and Ogren, 1983; Zhu et al., 2004), and there are some indications that the Sc/o of different organisms may be close to optimal for their respective environments (Tcherkez et al., 2006; Savir et al., 2010). Second, a CO2-concentrating mechanism could be engineered into C3 plants. For example, introducing cyanobacterial bicarbonate transporters (Price et al., 2011) or introducing C4 metabolism could be used to concentrate CO2 in the vicinity of Rubisco and, thereby, suppress the oxygenation reaction of Rubisco (Furbank and Hatch, 1987; Mitchell and Sheehy, 2006). Past efforts to introduce a C4 pathway into C3 plants have focused on biochemical reactions related to C4 photosynthesis without taking into account the anatomical differences between C3 and C4 plants, which may have been responsible for the limited success of such endeavors (Fukayama et al., 2003). Recently, there has been renewed interest in engineering C4 photosynthetic pathways into C3 plants, with efforts focusing on understanding and engineering the genetic regulatory network related to the control of both the anatomical and biochemical properties related to C4 photosynthesis (Mitchell and Sheehy, 2006; Langdale, 2011).Transgenic approaches have been used to knock down or knock out enzymes in the photorespiratory pathway. Unfortunately, the inhibition of photorespiration by the deletion or down-regulation of enzymes in the photorespiratory pathway resulted in a conditional lethal phenotype (i.e. such plants cannot survive under ambient oxygen and CO2 concentrations but may be rescued by growing them under low-oxygen or high-CO2 conditions; for review, see Somerville and Ogren, 1982; Somerville, 2001). Another approach to reduce photorespiration is to block (or inhibit) enzymes in this pathway using chemical inhibitors. Zelitch (1966, 1974, 1979) reported that net photosynthesis increased by inhibiting glycolate oxidase or glycolate synthesis. However, other groups showed that the inhibition of glycolate oxidase or Gly decarboxylation led to the inhibition of photosynthesis (Chollet, 1976; Kumarasinghe et al., 1977; Servaites and Ogren, 1977; Baumann et al., 1981). It turns out that plants cannot efficiently metabolize photorespiratory intermediates without a photorespiratory pathway, and suppression of this pathway inhibits the recycling of carbon back toward RuBP, which is necessary for maintaining the CBB cycle (Peterhansel et al., 2010; Peterhansel and Maurino, 2011). Moreover, the accumulation of toxic metabolic intermediates (e.g. P-Gly) can strongly inhibit photosynthesis (Anderson, 1971; Kelly and Latzko, 1976; Chastain and Ogren, 1989; Campbell and Ogren, 1990). This may explain why earlier attempts to block or reduce photorespiration have failed to improve carbon gain.Instead of reducing photorespiration directly, a promising idea is to engineer a photorespiratory bypass pathway. Such a pathway would metabolize P-Gly produced by RuBP oxygenation but minimize carbon, nitrogen, and energy losses and avoid the accumulation of photorespiratory intermediates. Kebeish et al. (2007) introduced the glycolate catabolic pathway from Escherichia coli into Arabidopsis (Arabidopsis thaliana); we will subsequently call this type of bypass the Kebeish bypass. In such transgenic plants, glycolate is converted to glycerate in the chloroplasts without ammonia release (Fig. 1). Previous studies suggested that this pathway theoretically requires less energy and shifts CO2 release from mitochondria to chloroplasts (Peterhansel and Maurino, 2011; Peterhansel et al., 2013); experimental results indicated that the bypass allowed for increased net photosynthesis and biomass production in Arabidopsis (Kebeish et al., 2007). There are reports of two other photorespiratory bypass pathways in the literature (Carvalho, 2005; Carvalho et al., 2011; Maier et al., 2012). In the Carvalho bypass (Carvalho, 2005; Carvalho et al., 2011), glyoxylate is converted to hydroxypyruvate in the peroxisome. Similar to the Kebeish bypass, the ammonia release is abolished, one-quarter of the carbon from glycolate is released as CO2 in the peroxisomes, and three-quarters of the carbon from glycolate is converted back to PGA. However, this pathway has only been partially realized in tobacco (Nicotiana tabacum); that is, the enzyme of the second reaction of this pathway was not detectable in the transgenic plants, and plants expressing this pathway showed stunted growth when grown in ambient air (Carvalho et al., 2011). The Maier bypass (Maier et al., 2012) is characterized by complete oxidation of glycolate in the chloroplasts. Initial results suggested that the photosynthesis and biomass of transgenic Arabidopsis with this pathway were enhanced (Maier et al., 2012).Recently, the design and benefits of the three bypass pathways were reviewed (Peterhansel et al., 2013), and it was suggested that a photorespiratory bypass can contribute to an enhanced photosynthetic CO2 uptake rate by lowering energy costs and minimizing carbon and nitrogen losses. However, a systematic and quantitative analysis of the potential contributions of these different factors to photosynthesis improvement has not yet been conducted. Systems modeling can help to design new metabolic pathways and improve our understanding of biochemical mechanisms (McNeil et al., 2000; Wendisch, 2005; Zhu et al., 2007; Bar-Even et al., 2010; Basler et al., 2012). Such models have been used successfully to gain insight into the photosynthetic metabolism (Laisk et al., 1989, 2006; Laisk and Edwards, 2000; Zhu et al., 2007, 2013; Wang et al., 2014). In this study, we use an extended kinetic model of C3 photosynthesis based on earlier work by Zhu et al. (2007) to systematically analyze the potential of three photorespiratory bypass pathways for improving photosynthetic efficiency (Supplemental Model S1). In addition, we determined under what conditions such bypass pathways may lead to increased photosynthesis and biomass production in C3 plants and how to further improve the photosynthesis of plants with such a bypass. Our analysis suggests that the benefit of a photorespiratory bypass varies dramatically if it is engineered into different crops.  相似文献   

4.
5.
Mixing of seawater subjects phytoplankton to fluctuations in photosynthetically active radiation (400–700 nm) and ultraviolet radiation (UVR; 280–400 nm). These irradiance fluctuations are now superimposed upon ocean acidification and thinning of the upper mixing layer through stratification, which alters mixing regimes. Therefore, we examined the photosynthetic carbon fixation and photochemical performance of a coccolithophore, Gephyrocapsa oceanica, grown under high, future (1,000 μatm) and low, current (390 μatm) CO2 levels, under regimes of fluctuating irradiances with or without UVR. Under both CO2 levels, fluctuating irradiances, as compared with constant irradiance, led to lower nonphotochemical quenching and less UVR-induced inhibition of carbon fixation and photosystem II electron transport. The cells grown under high CO2 showed a lower photosynthetic carbon fixation rate but lower nonphotochemical quenching and less ultraviolet B (280–315 nm)-induced inhibition. Ultraviolet A (315–400 nm) led to less enhancement of the photosynthetic carbon fixation in the high-CO2-grown cells under fluctuating irradiance. Our data suggest that ocean acidification and fast mixing or fluctuation of solar radiation will act synergistically to lower carbon fixation by G. oceanica, although ocean acidification may decrease ultraviolet B-related photochemical inhibition.The oceans absorb about 25 million tons of CO2 per day from the atmosphere (Sabine et al., 2004), leading to the acidification of seawater in surface oceans. The pH of oceanic surface seawater will decline by 0.3 to 0.4 units, reflecting a 100% to 150% increase in [H+], by the year 2100 under “a fossil-fuel intensive” emission scenario (Houghton, 2001). This ocean acidification and the associated chemical changes may bring about critical ecological and social consequences (Turley et al., 2010).Coccolithophores, as a key group of oceanic primary producers, with coccolith scales made of CaCO3, are important to global carbon cycles (Riebesell and Tortell, 2011). Ocean acidification generally decreases calcification by coccolithophores (Riebesell et al., 2000; Zondervan et al., 2002; Delille et al., 2005; Beaufort et al., 2011) and other calcifying algae (Gao and Zheng, 2010; Sinutok et al., 2011), with responses differing across species or different environmental conditions (Langer et al., 2006, 2009; Iglesias-Rodriguez et al., 2008; Doney et al., 2009; Shi et al., 2009). Algal calcification, in turn, influences the impacts of solar ultraviolet radiation (UVR; 280–400 nm) on the algae’s photophysiology (Gao et al., 2009; Gao and Zheng, 2010; Guan and Gao, 2010).Although the Montreal Protocol has resulted in a slowing of ozone depletion, ultraviolet B (UVB) irradiance (280–315 nm) reaching northern temperate regions increased 10% between 1983 and 2003 (Josefsson, 2006), and a recent observation found an ozone hole above the Arctic (Manney et al., 2011), reflecting ongoing impacts of climate change on ozone depletion. Biologically significant levels of UVR reach as deep as 80 m in pelagic oceans (Smith et al., 1992). In coastal waters or areas with high productivity, UVB irradiance usually penetrates only a few meters due to the attenuation caused by suspended particles and dissolved organic matter (Hargreaves, 2003; Tedetti and Sempéré, 2006). Ultraviolet A (UVA) and photosynthetically active radiation (PAR) are also attenuated but penetrate to much deeper depths due to their wavelength properties and intensities. UVA and UVB can both act synergistically with ocean acidification to inhibit algal photosynthetic performance (Gao et al., 2009), and the inhibition caused by UVB could be about 2.5 times that caused by UVA (Gao and Zheng, 2010); however, an antagonistic effect of UVB and ocean acidification was also found in a diatom (Li et al., 2012).In parallel, global warming due to increased atmospheric CO2 concentration causes ocean warming, which results in a decrease in the depth of the upper mixing layer (UML; Sarmiento et al., 2004). Such stratification increases integrated exposures of phytoplankton cells within the UML to solar UVR and visible radiation and decreases the upward transport of nutrients from deeper water layers, influencing phytoplankton photophysiology (Beardall et al., 2009; Gao et al., 2012a). Fluctuations of both solar PAR and UVR within the UML affect phytoplankton photosynthetic activity and carbon fixation (Helbling et al., 2003; Villafañe et al., 2007; Guan and Gao, 2008; Dimier et al., 2009). Mixing depths and/or mixing rates in the upper oceans also change in response to increased stratification and/or wind speed due to global climate change (Sarmiento et al., 2004; Boyd et al., 2010).Phytoplankton responses to fluctuating solar radiation vary, particularly if considered in combination with other environmental factors, due to antagonistic or synergistic interactions. Fluctuation of solar radiation on cloudy days led to higher primary production in the presence of UVA (315–400 nm) as compared with the presence of UVA on sunny days (Gao et al., 2007). Algal acclimation to fluctuating irradiance can lead to differences in growth rates and cellular pigment content compared with the cells acclimated to constant irradiance (Van de Poll et al., 2007, 2010). On the other hand, the mixing rate in the UML is strongly controlled by wind (Denman and Gargett, 1983; MacIntyre, 1993), which may increase due to global warming (Toggweiler and Russell, 2008). Therefore, changes in mixing rate and stratification may interact with ocean acidification to affect the photophysiology of phytoplankton. Nevertheless, to our knowledge, nothing has yet been documented on the combined impacts of the fluctuation of PAR or UVR and ocean acidification on the photosynthetic performance of coccolithophores.Under this scenario, we expect that the photosynthesis of coccolithophores will respond differentially to fluctuating PAR, with or without UVR, when grown under ocean acidification conditions, since the balance of high PAR- or UVR-induced damage and the counteracting repair could differ under elevated CO2 or acidity. To test these interactions, we grew Gephyrocapsa oceanica, which is widely distributed in temperate and tropical waters (Okabe, 1997), under current and ocean acidification conditions and examined its photochemical activity and photosynthesis under different combinations of fluctuating PAR and UVR.  相似文献   

6.
7.
8.
Vertical leaf nitrogen (N) gradient within a canopy is classically considered as a key adaptation to the local light environment that would tend to maximize canopy photosynthesis. We studied the vertical leaf N gradient with respect to the light gradient for wheat (Triticum aestivum) canopies with the aims of quantifying its modulation by crop N status and genetic variability and analyzing its ecophysiological determinants. The vertical distribution of leaf N and light was analyzed at anthesis for 16 cultivars grown in the field in two consecutive seasons under two levels of N. The N extinction coefficient with respect to light (b) varied with N supply and cultivar. Interestingly, a scaling relationship was observed between b and the size of the canopy for all the cultivars in the different environmental conditions. The scaling coefficient of the b-green area index relationship differed among cultivars, suggesting that cultivars could be more or less adapted to low-productivity environments. We conclude that the acclimation of the leaf N gradient to the light gradient is a whole-plant process that depends on canopy size. This study demonstrates that modeling leaf N distribution and canopy expansion based on the assumption that leaf N distribution parallels that of the light is inappropriate. We provide a robust relationship accounting for vertical leaf N gradient with respect to vertical light gradient as a function of canopy size.In cereals, as in many crop species, nitrogen (N) nutrition is a major determinant in the elaboration of grain yield and quality (Lemaire and Millard, 1999; Lawlor, 2002; Hikosaka, 2005). N is involved in both meristematic and photosynthetic activities, with consequences on plant architecture and carbon acquisition and in fine on grain yield and protein concentration. Beside the total amount of N absorbed by the crop, the allocation of N among plant organs plays a key role in determining crop productivity and quality (Grindlay, 1997; Dreccer et al., 1998; Hikosaka, 2005).Light interception and leaf N content are the two main factors governing carbon assimilation at the leaf scale (Evans, 1989). For various species, both light and leaf N attenuate with cumulative leaf area index counted from the top of the canopy (Field, 1983; Hirose and Werger, 1987). Leaf N vertical gradients have been regarded as an adaptive response to the local light environment, maximizing canopy photosynthesis and N utilization efficiency (Hirose and Werger, 1987; Hikosaka et al., 1994; Drouet and Bonhomme, 1999), as N is largely contained in the assimilatory enzyme Rubisco. Theoretical studies indicated that leaf N maximizes canopy photosynthesis when it parallels the light gradient (i.e. when the light [KL] and N [KN] extinction coefficients are equal), considering that the leaf N gradient is “optimal” in accordance with the “optimization theory” (Field, 1983; Hirose and Werger, 1987; Anten et al., 1995b).Factors other than the photosynthetic photon flux density (PPFD) might be responsible for the observed leaf N distribution. For instance, the acropetal gradients of leaf age (Hikosaka et al., 1994; Hikosaka, 2005) and light composition (Rousseaux et al., 1999) are known to strengthen the leaf N gradient. However, the impact of each of these factors has been shown to be much less than that of the PPFD gradient (Werger and Hirose, 1991; Pons and de Jong-van Berkel, 2004), although for the grass species Brachypodium pinnatum other factors than light might be involved (Pons et al., 1993). At the molecular level, the process could be driven by the import of compounds such as cytokinins transported in the transpiration stream (Pons et al., 2001; Boonman et al., 2007). Although the actual N distribution usually follows the light gradient, in all studies it is less steep than the calculated optimal N profile maximizing canopy photosynthesis (Pons et al., 1989; Yin et al., 2003). Possible reasons for this discrepancy have been discussed in detail by Kull (2002). Sink-source relations and in particular the demand for N could modulate the light-leaf N relationship (Dreccer et al., 1998), but conflicting results have been reported regarding the effect of N availability on the light-leaf N relationship. While some authors found no effect of N availability (Sinclair and Shiraiwa, 1993; Milroy et al., 2001), others found that the N gradient relative to light (i.e. KL/KN) was steeper under low N (Hikosaka et al., 1994; Grindlay et al., 1995; Lötscher et al., 2003) or that the response of the light-leaf N relationship to N availability depended on the developmental stage (Dreccer et al., 2000). Interspecific differences in the light-leaf N relationship have also been reported and were related to differences in phenotypic plasticity (Aerts, 1996) or plant architecture (leaf stature and branching pattern; Anten et al., 1995a; Lötscher et al., 2003).Since canopy photosynthesis is dependent upon the leaf N gradient, it has been suggested that the pattern of leaf N distribution could be responsible for part of the genetic variability associated with the negative correlation between grain yield and protein concentration reported for various crop species (Dreccer et al., 1998). In wheat (Triticum aestivum), N accumulated before anthesis contributes 30% to 70% of grain N (Mi et al., 2000; Kichey et al., 2007). The efficiency of N translocation from the lower to the upper leaves may increase with the steepness of the N gradient, with only a negligible effect on canopy carbon gain integrated over the whole grain-filling period. This hypothesis is consistent with experimental studies based on a range of genotypes showing that, at a given grain yield level, grain protein concentration is positively related to the efficiency of N translocation either from the lower to the upper leaves (Wang et al., 2005) or from the leaves to the grains (Monaghan et al., 2001; Jukanti and Fischer, 2008). Only a few studies have investigated the intraspecific variability of the light-N relationship at the intraspecific level (Shiraiwa and Sinclair, 1993; Bindraban, 1999; Bertheloot et al., 2008; van Oosterom et al., 2010). For wheat, published analyses of the genetic variability of the light-leaf N relationship were limited to only two to five genotypes, and no genetic differences were found (Bindraban, 1999; Bertheloot et al., 2008).This paper focuses on the genetic variability of the vertical leaf N gradient with respect to light for wheat. Three main issues were investigated. What is the effect of N supply on the vertical distribution of leaf N? Does the distribution of leaf N with respect to light differ among genotypes? If the adjustment of leaf N to the light gradient varies with both the genotype and N supply, could this genetic and environmental variability have a unique ecophysiological determinant (driving variable)?These questions were addressed using 16 genotypes (Supplemental Table S1) covering a wide range of variation for N use efficiency (i.e. grain dry mass yield per unit of available mineral N from the soil and fertilizer), for grain protein concentration (Le Gouis et al., 2000; Foulkes et al., 2006; Gaju et al., 2011) and for the deviation from the negative correlation between grain yield and protein concentration (Oury et al., 2003). The 16 genotypes were grown in the field under two conditions of N supply (N− and N+ for low- and high-N treatments, respectively) in order to modulate crop N status at Clermont-Ferrand (CF) in France in two consecutive seasons (experiments CF07 and CF08). In addition, four of the 16 cultivars representing the variability observed for N utilization and N uptake efficiency were grown in the field under two conditions of N supply at Sutton Bonington (SB) in the United Kingdom in one season (experiment SB07). The distribution of leaf N was analyzed at anthesis. The first reason for this is that the distribution of both light and leaf N within the canopy is relatively stable from this phenological stage until almost the end of grain filling (Bertheloot et al., 2008). Whereas the canopy green area index (GAI) decreases dramatically during the grain-filling period, the structure of the canopy affecting light interception does not change significantly during that period. Both the vertical light and N distributions down the canopy are unchanged during most of the grain-filling period; therefore, the KN-to-KL ratio is constant during that period (Bertheloot et al., 2008). Similarly, Archontoulis et al. (2011) showed that KN-to-KL ratio is not modified during the vegetative and reproductive stages for field-gown sunflower (Helianthus annuus) crops. Therefore, as most of the final grain yield results from carbon assimilated after anthesis (Bidinger et al., 1977; Gebbing and Schnyder, 1999), the N distribution at anthesis is very relevant in terms of carbon assimilation and grain yield in wheat. A second reason is that the number and potential size of grains are determined around anthesis, which therefore appears as a critical stage in the formation of grain yield. A better understanding of the ecophysiological determinants of leaf N gradient at this phenological stage could consequently be crucial for improving wheat productivity and quality (Dreccer et al., 1998).  相似文献   

9.
10.
In addition to the linear electron flow, a cyclic electron flow (CEF) around photosystem I occurs in chloroplasts. In CEF, electrons flow back from the donor site of photosystem I to the plastoquinone pool via two main routes: one that involves the Proton Gradient Regulation5 (PGR5)/PGRL1 complex (PGR) and one that is dependent of the NADH dehydrogenase-like complex. While the importance of CEF in photosynthesis and photoprotection has been clearly established, little is known about its regulation. We worked on the assumption of a redox regulation and surveyed the putative role of chloroplastic thioredoxins (TRX). Using Arabidopsis (Arabidopsis thaliana) mutants lacking different TRX isoforms, we demonstrated in vivo that TRXm4 specifically plays a role in the down-regulation of the NADH dehydrogenase-like complex-dependent plastoquinone reduction pathway. This result was confirmed in tobacco (Nicotiana tabacum) plants overexpressing the TRXm4 orthologous gene. In vitro assays performed with isolated chloroplasts and purified TRXm4 indicated that TRXm4 negatively controls the PGR pathway as well. The physiological significance of this regulation was investigated under steady-state photosynthesis and in the pgr5 mutant background. Lack of TRXm4 reversed the growth phenotype of the pgr5 mutant, but it did not compensate for the impaired photosynthesis and photoinhibition sensitivity. This suggests that the physiological role of TRXm4 occurs in vivo via a mechanism distinct from direct up-regulation of CEF.In plant thylakoids, photosynthesis involves a linear electron flow (LEF) from water to NADP+ via PSII, cytochrome b6/f, PSI, and soluble carriers. LEF produces NADPH and generates a transthylakoidal electrochemical proton gradient that drives the synthesis of ATP. Besides LEF, cyclic electron flow (CEF) can also occur, involving only PSI (for review, see Johnson, 2011; Kramer and Evans, 2011). These additional reactions include two main distinct pathways involving either the Proton Gradient Regulation5 (PGR5)/PGRL1 complex (Munekage et al., 2002; DalCorso et al., 2008) or the NADH dehydrogenase-like complex (NDH; for review, see Battchikova et al., 2011; Ifuku et al., 2011). The functioning of either CEF pathway, which generates a pH gradient ΔpH without any accumulation of NADPH, is thought to achieve the appropriate ATP/NADPH balance required for the biochemical needs of the plant, especially under certain environmental conditions such as low CO2 (Golding and Johnson, 2003), heat (Clarke and Johnson, 2001), cold (Clarke and Johnson, 2001), drought (Golding and Johnson, 2003; Kohzuma et al., 2009), high light (Munekage et al., 2004), or dark-to-light transitions (Joliot and Joliot, 2005; Fan et al., 2007). CEF-generated ΔpH is also involved in photoprotection owing to the down-regulation of PSII via nonphotochemical quenching (Munekage et al., 2004; Takahashi et al., 2009). Very recently, the role of the PGR5 protein as a regulator of LEF has been established. It has proved to be essential in the protection of PSI from photodamage (Suorsa et al., 2012).The two cyclic pathways are redundant (Munekage et al., 2004), sharing ferredoxin (Fd) as a common stromal electron donor (Yamamoto et al., 2011) and electron carriers from plastoquinone (PQ) to PSI with LEF. Thus, LEF and either of the CEF pathways may be in competition. The molecular events that allow CEF to challenge LEF remain enigmatic, particularly when considering that the conditions that require CEF are also those under which LEF is in excess. Efforts to understand the appropriate functioning of CEF have led to the proposition of several models segregating cyclic and linear pathways at a structural level (for review, see Eberhard et al., 2008; Cardol et al., 2011; Johnson, 2011; Rochaix, 2011). According to the restricted diffusion model, founded on the uneven distribution of the photosynthetic protein complexes in the thylakoids, there is little competition between CEF and LEF, as CEF occurs in stroma lamellae where PSI is concentrated while LEF takes place in the grana stacks. In line with the supercomplex model, whose relevance was demonstrated in the microalga Chlamydomonas reinhardtii, CEF happens within tightly bound supercomplexes containing PSI, with its own light-harvesting complex (LHCI), the PSII light-harvesting complex (LHCII), cytochrome b6/f, Fd, Fd NADP reductase (FNR), and the integral membrane protein PGRL1 (Iwai et al., 2010). In higher plants, an association between NDH and PSI subunits suggests the formation of such supercomplexes (Peng et al., 2009). The availability of FNR, found either free in the stroma or bound to the thylakoids (Zhang et al., 2001), has also been proposed to modulate partitioning between LEF and CEF (Joliot and Joliot, 2006; Joliot and Johnson, 2011). In addition, more dynamic models that illustrate competitive processes involved in the distribution of electrons between the cyclic and linear flows have been proposed. The competition between cytochrome b6/f and FNR for electrons from Fd could regulate the segregation between LEF and CEF (Breyton et al., 2006; Yamamoto et al., 2006; Hald et al., 2008). A few years ago, Joliot and Joliot (2006) suggested that the ATP/ADP ratio was one of the parameters that triggered on the transition between LEF and CEF. It was also established that the redox poise of chloroplast stroma contributed to the regulation of the photosynthetic pathway and played an important role in defining the extent of CEF. Breyton et al. (2006) scrutinized this redox regulation and established that the fraction of PSI complexes engaged in CEF could be modulated by changes in the stromal redox state. Overreduction of the NADPH pool was involved in the repartition between LEF and CEF (Joliot and Joliot, 2006). The NADPH/NADP+ ratio was proposed as a regulator of PGR-dependent CEF in vivo (Okegawa et al., 2008).All the published data supporting a role for the redox status in the regulation of CEF urged us to investigate a putative role of thioredoxins (TRX) in the regulation of CEF. TRX are ubiquitous disulfide reductases regulating the redox status of target proteins (for review, see Lemaire et al., 2007; Meyer et al., 2009). In chloroplast, TRX mediate the light regulation of numerous enzymes, among which some belong to the Calvin cycle (for review, see Schürmann and Buchanan, 2008; Montrichard et al., 2009; Lindahl et al., 2011). Global proteomic approaches have revealed that well-known photosynthetic complex subunits may be partners of TRX, such as PsbO in PSII, plastocyanin, Rieske Fe-S protein in cytochrome b6/f, and PsaK and PsaN in PSI (for review, see Montrichard et al., 2009; Lindahl et al., 2011). Furthermore, regarding the regulation of photosynthesis, TRX have also been involved in state transitions (Rintamäki et al., 2000; Buchanan and Balmer, 2005), and their participation in the control of the redox poise of the electron transport chain has also been suggested (Johnson, 2003).In this work, we have investigated the possible role of TRX in the regulation of CEF. Using Arabidopsis (Arabidopsis thaliana) mutants with altered expression of genes encoding different plastid TRX, we have established in vivo the inhibitor activity of TRXm4 on the NDH-dependent pathway for plastoquinone reduction. This result was confirmed in transplastomic tobacco (Nicotiana tabacum) plants overexpressing the TRXm4 orthologous gene. Moreover, in vitro assays performed with isolated chloroplasts indicated that TRXm4 negatively controls the PGR-dependent electron flow as well.  相似文献   

11.
Hydrogen sulfide (H2S) is the third biological gasotransmitter, and in animals, it affects many physiological processes by modulating ion channels. H2S has been reported to protect plants from oxidative stress in diverse physiological responses. H2S closes stomata, but the underlying mechanism remains elusive. Here, we report the selective inactivation of current carried by inward-rectifying K+ channels of tobacco (Nicotiana tabacum) guard cells and show its close parallel with stomatal closure evoked by submicromolar concentrations of H2S. Experiments to scavenge H2S suggested an effect that is separable from that of abscisic acid, which is associated with water stress. Thus, H2S seems to define a unique and unresolved signaling pathway that selectively targets inward-rectifying K+ channels.Hydrogen sulfide (H2S) is a small bioactive gas that has been known for centuries as an environmental pollutant (Reiffenstein et al., 1992). H2S is soluble in both polar and, especially, nonpolar solvents (Wang, 2002), and has recently come to be recognized as the third member of a group of so-called biological gasotransmitters. Most importantly, H2S shows both physical and functional similarities to the other gasotransmitters nitric oxide (NO) and carbon monoxide (Wang, 2002), and it has been shown to participate in diverse physiological processes in animals, including cardioprotection, neuromodulation, inflammation, apoptosis, and gastrointestinal functions among others (Kabil et al., 2014). Less is known about H2S molecular targets and its modes of action. H2S can directly modify specific targets through protein sulfhydration (the addition of an -SH group to thiol moiety of proteins; Mustafa et al., 2009) or reaction with metal centers (Li and Lancaster, 2013). It can also act indirectly, reacting with NO to form nitrosothiols (Whiteman et al., 2006; Li and Lancaster, 2013). Among its molecular targets, H2S has been reported to regulate ATP-dependent K+ channels (Yang et al., 2005), Ca2+-activated K+ channels, T- and L-type Ca2+ channels, and transient receptor potential channels (Tang et al., 2010; Peers et al., 2012), suggesting H2S as a key regulator of membrane ion transport.In plants, H2S is produced enzymatically by the desulfhydration of l-Cys to form H2S, pyruvate, and ammonia in a reaction catalyzed by the enzyme l-Cys desulfhydrase (Riemenschneider et al., 2005a, 2005b), DES1, that has been characterized in Arabidopsis (Arabidopsis thaliana; Alvarez et al., 2010). Alternatively, H2S can be produced from d-Cys by d-Cys desulfhydrase (Riemenschneider et al., 2005a, 2005b) and in cyanide metabolism by β-cyano-Ala synthase (García et al., 2010). H2S action was originally related to pathogenesis resistance (Bloem et al., 2004), but in the last decade it has been proven to have an active role in signaling, participating in key physiological processes, such as germination and root organogenesis (Zhang et al., 2008, 2009a), heat stress (Li et al., 2013a, 2013b), osmotic stress (Zhang et al., 2009b), and stomatal movement (García-Mata and Lamattina, 2010; Lisjak et al., 2010, 2011; Jin et al., 2013). Moreover, H2S was reported to participate in the signaling of plant hormones, including abscisic acid (ABA; García-Mata and Lamattina, 2010; Lisjak et al., 2010; Jin et al., 2013; Scuffi et al., 2014), ethylene (Hou et al., 2013), and auxin (Zhang et al., 2009a).ABA is an important player in plant physiology. Notably, upon water stress, ABA triggers a complex signaling network to restrict the loss of water through the transpiration stream, balancing these needs with those of CO2 for carbon assimilation. In the guard cells that surround the stomatal pore, ABA induces an increase of cytosolic-free Ca2+ concentration ([Ca2+]cyt), elevates cytosolic pH (pHi), and activates the efflux of anions, mainly chloride, through S- and R-type anion channels. The increase in [Ca2+]cyt inactivates inward-rectifying K+ channels (IKIN); anion efflux depolarizes the plasma membrane, and together with the rise in pHi, it activates K+ efflux through outward-rectifying K+ channels (IKOUT; Blatt, 2000; Schroeder et al., 2001). These changes in ion flux, in turn, generate an osmotically driven reduction in turgor and volume and closure of the stomatal pore. All three gasotransmitters have been implicated in regulating the activity of guard cell ion channels, but direct evidence is available only for NO (Garcia-Mata et al., 2003; Sokolovski et al., 2005). Here, we have used two-electrode voltage clamp measurements to study the role of H2S in the regulation of the guard cell K+ channels of tobacco (Nicotiana tabacum). Our results show that H2S selectively inactivates IKIN and that this action parallels that of stomatal closure. These results confirm H2S as a unique factor regulating guard cell ion transport and indicate that H2S acts in a manner separable from that of ABA.  相似文献   

12.
The multifunctional movement protein (MP) of Tomato mosaic tobamovirus (ToMV) is involved in viral cell-to-cell movement, symptom development, and resistance gene recognition. However, it remains to be elucidated how ToMV MP plays such diverse roles in plants. Here, we show that ToMV MP interacts with the Rubisco small subunit (RbCS) of Nicotiana benthamiana in vitro and in vivo. In susceptible N. benthamiana plants, silencing of NbRbCS enabled ToMV to induce necrosis in inoculated leaves, thus enhancing virus local infectivity. However, the development of systemic viral symptoms was delayed. In transgenic N. benthamiana plants harboring Tobacco mosaic virus resistance-22 (Tm-22), which mediates extreme resistance to ToMV, silencing of NbRbCS compromised Tm-22-dependent resistance. ToMV was able to establish efficient local infection but was not able to move systemically. These findings suggest that NbRbCS plays a vital role in tobamovirus movement and plant antiviral defenses.Plant viruses use at least one movement protein (MP) to facilitate viral spread between plant cells via plasmodesmata (PD; Lucas and Gilbertson, 1994; Ghoshroy et al., 1997). Among viral MPs, the MP of tobamoviruses, such as Tobacco mosaic virus (TMV) and its close relative Tomato mosaic virus (ToMV), is the best characterized. TMV MP specifically accumulates in PD and modifies the plasmodesmatal size exclusion limit in mature source leaves or tissues (Wolf et al., 1989; Deom et al., 1990; Ding et al., 1992). TMV MP and viral genomic RNA form a mobile ribonucleoprotein complex that is essential for cell-to-cell movement of viral infection (Watanabe et al., 1984; Deom et al., 1987; Citovsky et al., 1990, 1992; Kiselyova et al., 2001; Kawakami et al., 2004; Waigmann et al., 2007). TMV MP also enhances intercellular RNA silencing (Vogler et al., 2008) and affects viral symptom development, host range, and host susceptibility to virus (Dardick et al., 2000; Bazzini et al., 2007). Furthermore, ToMV MP is identified as an avirulence factor that is recognized by tomato (Solanum lycopersicum) resistance proteins Tobacco mosaic virus resistance-2 (Tm-2) and Tm-22 (Meshi et al., 1989; Lanfermeijer et al., 2004). Indeed, tomato Tm-22 confers extreme resistance against TMV and ToMV in tomato plants and even in heterologous tobacco (Nicotiana tabacum) plants (Lanfermeijer et al., 2003, 2004).To date, several host factors that interact with TMV MP have been identified. These TMV MP-binding host factors include cell wall-associated proteins such as pectin methylesterase (Chen et al., 2000), calreticulin (Meshi et al., 1989), ANK1 (Ueki et al., 2010), and the cellular DnaJ-like protein MPIP1 (Shimizu et al., 2009). Many cytoskeletal components such as actin filaments (McLean et al., 1995), microtubules (Heinlein et al., 1995), and the microtubule-associated proteins MPB2C (Kragler et al., 2003) and EB1a (Brandner et al., 2008) also interact with TMV MP. Most of these factors are involved in TMV cell-to-cell movement.Rubisco catalyzes the first step of CO2 assimilation in photosynthesis and photorespiration. The Rubisco holoenzyme is a heteropolymer consisting of eight large subunits (RbCLs) and eight small subunits (RbCSs). RbCL was reported to interact with the coat protein of Potato virus Y (Feki et al., 2005). Both RbCS and RbCL were reported to interact with the P3 proteins encoded by several potyviruses, including Shallot yellow stripe virus, Onion yellow dwarf virus, Soybean mosaic virus, and Turnip mosaic virus (Lin et al., 2011). Proteomic analysis of the plant-virus interactome revealed that RbCS participates in the formation of virus complexes of Rice yellow mottle virus (Brizard et al., 2006). However, the biological function of Rubisco in viral infection remains unknown.In this study, we show that RbCS plays an essential role in virus movement, host susceptibility, and Tm-22-mediated extreme resistance in the ToMV-host plant interaction.  相似文献   

13.
14.
A major contributor to the global carbon cycle is plant respiration. Elevated atmospheric CO2 concentrations may either accelerate or decelerate plant respiration for reasons that have been uncertain. We recently established that elevated CO2 during the daytime decreases plant mitochondrial respiration in the light and protein concentration because CO2 slows the daytime conversion of nitrate (NO3) into protein. This derives in part from the inhibitory effect of CO2 on photorespiration and the dependence of shoot NO3 assimilation on photorespiration. Elevated CO2 also inhibits the translocation of nitrite into the chloroplast, a response that influences shoot NO3 assimilation during both day and night. Here, we exposed Arabidopsis (Arabidopsis thaliana) and wheat (Triticum aestivum) plants to daytime or nighttime elevated CO2 and supplied them with NO3 or ammonium as a sole nitrogen (N) source. Six independent measures (plant biomass, shoot NO3, shoot organic N, 15N isotope fractionation, 15NO3 assimilation, and the ratio of shoot CO2 evolution to O2 consumption) indicated that elevated CO2 at night slowed NO3 assimilation and thus decreased dark respiration in the plants reliant on NO3. These results provide a straightforward explanation for the diverse responses of plants to elevated CO2 at night and suggest that soil N source will have an increasing influence on the capacity of plants to mitigate human greenhouse gas emissions.The CO2 concentration in Earth’s atmosphere has increased from about 270 to 400 µmol mol–1 since 1800, and may double before the end of the century (Intergovernmental Panel on Climate Change, 2013). Plant responses to such increases are highly variable, but plant nitrogen (N) concentrations generally decline under elevated CO2 (Cotrufo et al., 1998; Long et al., 2004). One explanation for this decline is that CO2 inhibits nitrate (NO3) assimilation into protein in the shoots of C3 plants during the daytime (Bloom et al., 2002, 2010, 2012, 2014; Cheng et al., 2012; Pleijel and Uddling, 2012; Myers et al., 2014; Easlon et al., 2015; Pleijel and Högy, 2015). This derives in part from the inhibitory effect of CO2 on photorespiration (Foyer et al., 2009) and the dependence of shoot NO3 assimilation on photorespiration (Rachmilevitch et al., 2004; Bloom, 2015).A key factor in global carbon budgets is plant respiration at night (Amthor, 1991; Farrar and Williams, 1991; Drake et al., 1999; Leakey et al., 2009). Nighttime elevated CO2 may inhibit, have a negligible effect on, or stimulate dark respiration, depending on the plant species (Bunce, 2001, 2003; Wang and Curtis, 2002), plant development stage (Wang et al., 2001; Li et al., 2013), experimental approach (Griffin et al., 1999; Baker et al., 2000; Hamilton et al., 2001; Bruhn et al., 2002; Jahnke and Krewitt, 2002; Bunce, 2004), and total N supply (Markelz et al., 2014). The current study is, to our knowledge, the first to examine the influence of N source, NO3 versus ammonium (NH4+), on plant dark respiration at elevated CO2 during the night.Plant organic N compounds account for less than 5% of the total dry weight of a plant, but conversion of NO3 into organic N expends about 25% of the total energy in shoots (Bloom et al., 1989) and roots (Bloom et al., 1992). During the day, photorespiration supplies a portion of the energy (Rachmilevitch et al., 2004; Foyer et al., 2009), but at night, this energetic cost is borne entirely by the respiration of C substrates (Amthor, 1995) and may divert a substantial amount of reductant from the mitochondrial electron transport chain (Cousins and Bloom, 2004). The relative importance of NO3 assimilation at night versus the day, however, is still a matter of intense debate (Nunes-Nesi et al., 2010). Here, we estimated NO3 assimilation using several independent methods and show in Arabidopsis (Arabidopsis thaliana) and wheat (Triticum aestivum), two diverse C3 plants, that NO3 assimilation at night can be substantial, and that elevated CO2 at night inhibits this process.  相似文献   

15.
Recently, a feedback inhibition of the chloroplastic 1-deoxy-d-xylulose 5-phosphate (DXP)/2-C-methyl-d-erythritol 4-phosphate (MEP) pathway of isoprenoid synthesis by end products dimethylallyl diphosphate (DMADP) and isopentenyl diphosphate (IDP) was postulated, but the extent to which DMADP and IDP can build up is not known. We used bisphosphonate inhibitors, alendronate and zoledronate, that inhibit the consumption of DMADP and IDP by prenyltransferases to gain insight into the extent of end product accumulation and possible feedback inhibition in isoprene-emitting hybrid aspen (Populus tremula × Populus tremuloides). A kinetic method based on dark release of isoprene emission at the expense of substrate pools accumulated in light was used to estimate the in vivo pool sizes of DMADP and upstream metabolites. Feeding with fosmidomycin, an inhibitor of DXP reductoisomerase, alone or in combination with bisphosphonates was used to inhibit carbon input into DXP/MEP pathway or both input and output. We observed a major increase in pathway intermediates, 3- to 4-fold, upstream of DMADP in bisphosphonate-inhibited leaves, but the DMADP pool was enhanced much less, 1.3- to 1.5-fold. In combined fosmidomycin/bisphosphonate treatment, pathway intermediates accumulated, reflecting cytosolic flux of intermediates that can be important under strong metabolic pull in physiological conditions. The data suggested that metabolites accumulated upstream of DMADP consist of phosphorylated intermediates and IDP. Slow conversion of the huge pools of intermediates to DMADP was limited by reductive energy supply. These data indicate that the DXP/MEP pathway is extremely elastic, and the presence of a significant pool of phosphorylated intermediates provides an important valve for fine tuning the pathway flux.Isoprenoids constitute a versatile class of compounds fulfilling major physiological functions. They are formed by two pathways in plants, the mevalonate (MVA) pathway in the cytosol (Gershenzon and Croteau, 1993) and the 1-deoxy-d-xylulose 5-phosphate (DXP)/2-C-methyl-d-erythritol 4-phosphate (MEP) pathway in plastids (Gershenzon and Croteau, 1993; Jomaa et al., 1999; Li and Sharkey, 2013b). The MVA pathway is primarily responsible for the synthesis of sesquiterpenes (C15), triterpenes (C30) including brassinosteroids, and even larger molecules such as dolichols (Bick and Lange, 2003; Li and Sharkey, 2013b; Rajabi Memari et al., 2013; Rosenkranz and Schnitzler, 2013). The DXP/MEP pathway is responsible for the synthesis of the simplest isoprenoids, isoprene and 2-methyl-3-buten-2-ol (C5), monoterpenes (C10), diterpenes (C20) including gibberellins and phytol residue of chlorophylls, and tetraterpenes (C40) including carotenoids (Rodríguez-Concepción and Boronat, 2002; Roberts, 2007).Given that in plants both pathways produce ultimately the same substrates, dimethylallyl diphosphate (DMADP) and isopentenyl diphosphate (IDP), the pertinent question is to what extent the two pathways can exchange metabolites (Rodríguez-Concepción and Boronat, 2002). There is evidence of a certain exchange of IDP between cytosolic and plastidic compartments, although the contribution of IDP from one compartment to the pathway flux in the other seems to be relatively minor (Schwender et al., 2001; Rodríguez-Concepción and Boronat, 2002; Bick and Lange, 2003). Some studies have further demonstrated that the exchange of IDP is fully bidirectional (De-Eknamkul and Potduang, 2003; Rodríguez-Concepción, 2006), whereas other studies suggest that IDP export from plastids to cytosol operates with a greater efficiency than the opposite transport (Hemmerlin et al., 2003; Laule et al., 2003). However, although the overall intercompartmental exchange of isoprenoid substrates to pathway flux in the given compartment might seem minor under nonstressed conditions, the importance of cross talk among the pathways might increase under stress conditions that specifically inhibit isoprenoid synthesis in one pathway. In fact, the DXP/MEP pathway is strongly linked to photosynthetic metabolism, and therefore, inhibition of photosynthesis under stressful conditions such as heat stress or drought or photoinhibition could inhibit the synthesis of isoprenoids when they are most needed to fulfill their protective function (Loreto and Schnitzler, 2010; Niinemets, 2010; Possell and Loreto, 2013). There is some evidence demonstrating a certain cooperativity among the two isoprenoid synthesis pathways under conditions leading to a reduction of the activity of one of them (Piel et al., 1998; Jux et al., 2001; Page et al., 2004; Rodríguez-Concepción, 2006), but the capacity for such a replacement of function and regulation is poorly understood.Recent studies using genetically modified plants accumulating end products of the DXP/MEP pathway or using natural variation in product accumulation have demonstrated the existence of a potentially important feedback regulation of the DXP/MEP pathway flux by primary end products of the pathway (Banerjee et al., 2013; Ghirardo et al., 2014; Wright et al., 2014). In particular, binding of DMADP and perhaps IDP to DXP synthase, the first enzyme in the DXP/MEP pathway, leads to downregulation of the pathway flux when the end products cannot be used, such as under stress conditions. However, the strength of such a feedback regulation can be importantly modified by accumulation of phosphorylated intermediates of the pathway, such that DMADP and IDP do not accumulate. Previous studies have demonstrated that there is a certain pool of phosphorylated intermediates in vivo, and that this pool can strongly increase under certain conditions, including experimental and genetic modification of DXP/MEP pathway input and output (Li et al., 2011; Rasulov et al., 2011; Li and Sharkey, 2013a; Ghirardo et al., 2014; Wright et al., 2014).It has further been shown that 2-C-methyl-d-erythritol 2,4-cyclodiphosphate (ME-cDP) is the metabolite accumulating in the plastids, and this accumulation can buffer DMADP and IDP changes in the case of varying DXP/MEP pathway input and consumption (Li and Sharkey, 2013a; Wright et al., 2014). A significant part of ME-cDP might even escape to cytosol, implying the existence of another interesting link between cytosolic and chloroplastic processes in isoprenoid synthesis (Wright et al., 2014). Furthermore, as ME-cDP is an important signaling molecule eliciting a number of gene expression responses (Xiao et al., 2012), accumulation of ME-cDP in plastids and flux to cytosol and further to the nucleus is particularly interesting from the perspective of long-term regulation of isoprenoid synthesis, and suggests a coordination of cellular stress responses by the plastidial isoprenoid synthesis pathway.Isoprene-emitting species constitute an exciting model system where a very large DXP/MEP pathway flux goes to isoprene synthesis under physiological conditions (Li and Sharkey, 2013b; Sharkey et al., 2013). In isoprene-emitting species, there is a concomitant use of the primary substrate DMADP between the plastidic synthesis of isoprene and isoprenoids with a larger molecular size, such as phytol residue of chlorophyll (C20) and carotenoids (C40; Ghirardo et al., 2014; Rasulov et al., 2014), and different from nonemitting species, isoprene emitters seem to support a much larger pool of DMADP without the onset of feedback inhibition (Ghirardo et al., 2014; Wright et al., 2014). However, it is poorly understood how inhibition of one branch of the pathway (isoprene versus larger isoprenoids) affects the other, to what extent it can lead to accumulation of phosphorylated intermediates, how it affects the overall pathway flux through the feedback regulation, and what is the possible role of cytosolic import and export of intermediates. These are all relevant questions to gain insight into the control of the partitioning of pathway flux between isoprene and larger isoprenoids and to understand the biological role of isoprene emission.Studies using metabolic inhibitors to deconvolute the factors involved in pathway regulation and understand the biological role of isoprene have so far used inhibitors that block the early steps of the corresponding pathways. In particular, fosmidomycin, a specific inhibitor of DXP reductoisomerase, the enzyme responsible for the synthesis of MEP from DXP, has been used to inhibit the DXP/MEP pathway (Loreto and Velikova, 2001; Sharkey et al., 2001; Loreto et al., 2004). In addition, lovastatin (mevinolin), the inhibitor of 3-hydroxy-3-methylglutaryl-CoA reductase that controls the MVA pathway flux, has been used to study the cooperativity of the two pathways (e.g. Laule et al., 2003; Mansouri and Salari, 2014). However, these inhibitors are not suitable to understand how end product accumulation can alter the pathway flux.Bisphosphonates constitute a promising class of inhibitors that could be particularly apt for studies on the effects of the inhibition of the end points of the pathway. They have been demonstrated to inhibit cytosolic farnesyl diphosphate (FDP) synthase activity (Oberhauser et al., 1998; Cromartie et al., 1999; van Beek et al., 1999; Bergstrom et al., 2000; Burke et al., 2004), аs well as geranyl diphosphate (GDP) and geranylgeranyl diphosphate (GGDP) synthase activities (Oberhauser et al., 1998; Cromartie et al., 1999; Kloer et al., 2006; No et al., 2012; Lindert et al., 2013). To our knowledge, bisphosphonates have not been used to study the effects of end product accumulation on the pathway flux in isoprene-emitting species, with the exception of one study that investigated the development of isoprene emission capacity through leaf ontogeny (Rasulov et al., 2014).A limitation with any inhibitor study could be a certain nonspecificity, inhibition of additional nondesired reactions, but so far there are no data on such nonspecificity of bisphosphonates. However, there is evidence that diphosphate and its analogs are inhibitors of any ferredoxin (Fd)-dependent reaction (Forti and Meyer, 1969; Bojko and Więckowski, 1999). This could be potentially relevant given that DXP/MEP pathway-reducing steps, at the level of 4-hydroxy-3-methyl-2-(E)-butenyl diphosphate (HMBDP) synthase (HDS) and reductase (HDR), directly accept electrons from Fd in light (Eisenreich et al., 2001; Seemann et al., 2006; Li and Sharkey, 2013a). In addition to the DXP/MEP pathway, inhibition at the level of Fd could also affect photosynthetic reactions and thereby alter energy supply for the DXP/MEP pathway.In this study, we have investigated the effects of inhibition of the initial and final steps of the DXP/MEP pathway by fosmidomycin and bisphosphonate inhibitors alendronate and zoledronate in a strong isoprene emitter hybrid aspen (Populus tremula × Populus tremuloides). Alendronate is a highly specific inhibitor of GDP (Lange et al., 2001; Burke et al., 2004) and FDP synthases (Bergstrom et al., 2000; Burke et al., 2004), and a less specific inhibitor of GGDP synthase (Szabo et al., 2002). Zoledronate operates similarly to alendronate, but is a much stronger inhibitor, being operationally active in concentrations several orders of magnitude less than alendronate (Lange et al., 2001; Henneman et al., 2011; Wasko, 2011). A unique in vivo method was used to study dynamic changes in DMADP and phosphorylated intermediate pool sizes (Rasulov et al., 2009a, 2011; Li et al., 2011), and different inhibitors were applied alone or in sequence to study the regulation of the pathway flux in conditions when the flux out of the pathway or into the pathway is curbed and when both the input and the output are curbed. Dynamic model calculations were used to quantitatively evaluate the significance of the cytosolic intermediate input into chloroplastic isoprenoid synthesis under different conditions of the DXP/MEP pathway entrance and exit, and to evaluate the possible nonspecific inhibition of other steps controlling DXP/MEP pathway flux. The study demonstrates the important regulation of DXP/MEP pathway input and output under conditions of end product accumulation and partial cooperativity among chloroplastic and cytosolic isoprenoid synthesis pathways.  相似文献   

16.
17.
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   

18.
19.
20.
The signaling role of hydrogen gas (H2) has attracted increasing attention from animals to plants. However, the physiological significance and molecular mechanism of H2 in drought tolerance are still largely unexplored. In this article, we report that abscisic acid (ABA) induced stomatal closure in Arabidopsis (Arabidopsis thaliana) by triggering intracellular signaling events involving H2, reactive oxygen species (ROS), nitric oxide (NO), and the guard cell outward-rectifying K+ channel (GORK). ABA elicited a rapid and sustained H2 release and production in Arabidopsis. Exogenous hydrogen-rich water (HRW) effectively led to an increase of intracellular H2 production, a reduction in the stomatal aperture, and enhanced drought tolerance. Subsequent results revealed that HRW stimulated significant inductions of NO and ROS synthesis associated with stomatal closure in the wild type, which were individually abolished in the nitric reductase mutant nitrate reductase1/2 (nia1/2) or the NADPH oxidase-deficient mutant rbohF (for respiratory burst oxidase homolog). Furthermore, we demonstrate that the HRW-promoted NO generation is dependent on ROS production. The rbohF mutant had impaired NO synthesis and stomatal closure in response to HRW, while these changes were rescued by exogenous application of NO. In addition, both HRW and hydrogen peroxide failed to induce NO production or stomatal closure in the nia1/2 mutant, while HRW-promoted ROS accumulation was not impaired. In the GORK-null mutant, stomatal closure induced by ABA, HRW, NO, or hydrogen peroxide was partially suppressed. Together, these results define a main branch of H2-regulated stomatal movement involved in the ABA signaling cascade in which RbohF-dependent ROS and nitric reductase-associated NO production, and subsequent GORK activation, were causally involved.Stomata are responsible for leaves of terrestrial plants taking in carbon dioxide for photosynthesis and likewise regulate how much water plants evaporate through the stomatal pores (Chaerle et al., 2005). When experiencing water-deficient conditions, surviving plants balance photosynthesis with controlling water loss through the stomatal pores, which relies on turgor changes by pairs of highly differentiated epidermal cells surrounding the stomatal pore, called the guard cells (Haworth et al., 2011; Loutfy et al., 2012).Besides the characterization of the significant roles of abscisic acid (ABA) in regulating stomatal movement, the key factors in guard cell signal transduction have been intensively investigated by performing forward and reverse genetics approaches. For example, both reactive oxygen species (ROS) and nitric oxide (NO) have been identified as vital intermediates in guard cell ABA signaling (Bright et al., 2006; Yan et al., 2007; Suzuki et al., 2011; Hao et al., 2012). The key ROS-producing enzymes in Arabidopsis (Arabidopsis thaliana) guard cells are the respiratory burst oxidase homologs (Rboh) D and F (Kwak et al., 2003; Bright et al., 2006; Mazars et al., 2010; Marino et al., 2012). Current available data suggest that there are at least two distinct pathways responsible for NO synthesis involved in ABA signaling in guard cells: the nitrite reductase (NR)- and l-Arg-dependent pathways (Desikan et al., 2002; Besson-Bard et al., 2008). Genetic evidence further demonstrated that removal of the major known sources of either ROS or NO significantly impairs ABA-induced stomatal closure. ABA fails to induce ROS production in the atrbohD/F double mutant (Kwak et al., 2003; Wang et al., 2012) and NO synthesis in the NR-deficient mutant nitrate reductase1/2 (nia1/2; Bright et al., 2006; Neill et al., 2008), both of which lead to impaired stomatal closure in Arabidopsis. Most importantly, ROS and NO, which function both synergistically and independently, have been established as ubiquitous signal transduction components to control a diverse range of physiological pathways in higher plants (Bright et al., 2006; Tossi et al., 2012).The guard cell outward-rectifying K+ channel (GORK) encodes the exclusive voltage-gated outwardly rectifying K+ channel protein, which was located in the guard cell membrane (Ache et al., 2000; Dreyer and Blatt, 2009). Expression profiles revealed that this gene is up-regulated upon the onset of drought, salinity, and cold stress and ABA exposure (Becker et al., 2003; Tran et al., 2013). Reverse genetic evidence further showed that GORK plays an important role in the control of stomatal movements and allows the plant to reduce transpirational water loss significantly (Hosy et al., 2003) and participates in the regulation of salinity tolerance by preventing salt-induced K+ loss (Jayakannan et al., 2013). Due to the high complexity of guard cell signaling cascades, whether and how ABA-triggered GORK up-regulation is attributed to the generation of cellular secondary messengers, such as ROS and NO, is less clear.Hydrogen gas (H2) was recently revealed as a signaling modulator with multiple biological functions in clinical trails (Ohsawa et al., 2007; Itoh et al., 2009; Ito et al., 2012). It was previously found that a hydrogenase system could generate H2 in bacteria and green algae (Meyer, 2007; Esquível et al., 2011). Although some earlier studies discovered the evolution of H2 in several higher plant species (Renwick et al., 1964; Torres et al., 1984), it was also proposed that the eukaryotic hydrogenase-like protein does not metabolize H2 (Cavazza et al., 2008; Mondy et al., 2014). Since the explosion limit of H2 gas is about 4% to 72.4% (v/v, in the air), the direct application of H2 gas in experiments is flammable and dangerous. Regardless of these problems to be resolved, the methodology, such as using exogenous hydrogen-rich water (HRW) or hydrogen-rich saline, which is safe, economical, and easily available, provides a valuable approach to investigate the physiological function of H2 in animal research and clinical trials. For example, hydrogen dissolved in Dulbecco’s modified Eagle’s medium was found to react with cytotoxic ROS and thus protect against oxidative damage in PC12 cells and rats (Ohsawa et al., 2007). The neuroprotective effect of H2-loaded eye drops on retinal ischemia-reperfusion injury was also reported (Oharazawa et al., 2010). In plants, corresponding results by using HRW combined with gas chromatography (GC) revealed that H2 could act as a novel beneficial gaseous molecule in plant responses against salinity (Xie et al., 2012; Xu et al., 2013), cadmium stress (Cui et al., 2013), and paraquat toxicity (Jin et al., 2013). More recently, the observation that HRW could delay the postharvest ripening and senescence of kiwifruit (Actinidia deliciosa) was reported (Hu et al., 2014).Considering the fact that the signaling cascades for salt, osmotic, and drought stresses share a common cascade in an ABA-dependent pathway, it would be noteworthy to identify whether and how H2 regulates the bioactivity of ABA-induced downstream components and, thereafter, biological responses, including stomatal closure and drought tolerance. To resolve these scientific questions, rbohD, rbohF, nia1/2, nitric oxide associated1 (noa1; Van Ree et al., 2011), nia1/2/noa1, and gork mutants were utilized to investigate the relationship among H2, ROS, NO, and GORK in the guard cell signal transduction network. By the combination of pharmacological and biochemical analyses with this genetics-based approach, we provide comprehensive evidence to show that H2 might be a newly identified bioeffective modulator involved in ABA signaling responsible for drought tolerance, that HRW-promoted stomatal closure was mainly attributed to the modulation of ROS-dependent NO generation, and that GORK might be the downstream target protein of H2 signaling.  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