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 31 毫秒
1.
2.
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   

3.
While interspecific variation in the temperature response of photosynthesis is well documented, the underlying physiological mechanisms remain unknown. Moreover, mechanisms related to species-dependent differences in photosynthetic temperature acclimation are unclear. We compared photosynthetic temperature acclimation in 11 crop species differing in their cold tolerance, which were grown at 15°C or 30°C. Cold-tolerant species exhibited a large decrease in optimum temperature for the photosynthetic rate at 360 μL L−1 CO2 concentration [Opt (A360)] when growth temperature decreased from 30°C to 15°C, whereas cold-sensitive species were less plastic in Opt (A360). Analysis using the C3 photosynthesis model shows that the limiting step of A360 at the optimum temperature differed between cold-tolerant and cold-sensitive species; ribulose 1,5-bisphosphate carboxylation rate was limiting in cold-tolerant species, while ribulose 1,5-bisphosphate regeneration rate was limiting in cold-sensitive species. Alterations in parameters related to photosynthetic temperature acclimation, including the limiting step of A360, leaf nitrogen, and Rubisco contents, were more plastic to growth temperature in cold-tolerant species than in cold-sensitive species. These plastic alterations contributed to the noted growth temperature-dependent changes in Opt (A360) in cold-tolerant species. Consequently, cold-tolerant species were able to maintain high A360 at 15°C or 30°C, whereas cold-sensitive species were not. We conclude that differences in the plasticity of photosynthetic parameters with respect to growth temperature were responsible for the noted interspecific differences in photosynthetic temperature acclimation between cold-tolerant and cold-sensitive species.The temperature dependence of leaf photosynthetic rate shows considerable variation between plant species and with growth temperature (Berry and Björkman, 1980; Cunningham and Read, 2002; Hikosaka et al., 2006). Plants native to low-temperature environments and those grown at low temperatures generally exhibit higher photosynthetic rates at low temperatures and lower optimum temperatures, compared with plants native to high-temperature environments and those grown at high temperatures (Mooney and Billings, 1961; Slatyer, 1977; Berry and Björkman, 1980; Sage, 2002; Salvucci and Crafts-Brandner, 2004b). For example, the optimum temperature for photosynthesis differs between temperate evergreen species and tropical evergreen species (Hill et al., 1988; Read, 1990; Cunningham and Read, 2002). Such differences have been observed even among ecotypes of the same species (Björkman et al., 1975; Pearcy, 1977; Slatyer, 1977).Temperature dependence of the photosynthetic rate has been analyzed using the biochemical model proposed by Farquhar et al. (1980). This model assumes that the photosynthetic rate (A) is limited by either ribulose 1,5-bisphosphate (RuBP) carboxylation (Ac) or RuBP regeneration (Ar). The optimum temperature for photosynthetic rate in C3 plants is thus potentially determined by (1) the temperature dependence of Ac, (2) the temperature dependence of Ar, or (3) both, at the colimitation point of Ac and Ar (Fig. 1; Farquhar and von Caemmerer, 1982; Hikosaka et al., 2006).Open in a separate windowFigure 1.A scheme illustrating the shift in the optimum temperature for photosynthesis depending on growth temperature. Based on the C3 photosynthesis model, the A360 (white and black circles) is limited by Ac (solid line) or Ar (broken line). The optimum temperature for the photosynthetic rate is potentially determined by temperature dependence of Ac (A), temperature dependence of Ar (B), or the intersection of the temperature dependences of Ac and Ar (C). When the optimum temperature for the photosynthetic rate shifts to a higher temperature, there are also three possibilities determining the optimum temperature: temperature dependence of Ac (D), temperature dependence of Ar (E), or the intersection of the temperature dependences of Ac and Ar (F). Especially in the case that the optimum temperature is determined by the intersection of the temperature dependences of Ac and Ar, the optimum temperature can shift by changes in the balance between Ac and Ar even when the optimum temperatures for these two partial reactions do not change.In many cases, the photosynthetic rate around the optimum temperature is limited by Ac, and thus the temperature dependence of Ac determines the optimum temperature for the photosynthetic rate (Hikosaka et al., 1999, 2006; Yamori et al., 2005, 2006a, 2006b, 2008; Sage and Kubien, 2007; Sage et al., 2008). As the temperature increases above the optimum, Ac is decreased by increases in photorespiration (Berry and Björkman, 1980; Jordan and Ogren, 1984; von Caemmerer, 2000). Furthermore, it has been suggested that the heat-induced deactivation of Rubisco is involved in the decrease in Ac at high temperature (Law and Crafts-Brandner, 1999; Crafts-Brandner and Salvucci, 2000; Salvucci and Crafts-Brandner, 2004a; Yamori et al., 2006b). Numerous previous studies have shown changes in the temperature dependence of Ac with growth temperature (Hikosaka et al., 1999; Bunce, 2000; Yamori et al., 2005). Also, the temperature sensitivity of Rubisco deactivation may differ between plant species (Salvucci and Crafts-Brandner, 2004b) and with growth temperature (Yamori et al., 2006b), which may explain variation in the optimum temperature for photosynthesis (Fig. 1, A and D).Ar is more responsive to temperature than Ac and often limits photosynthesis at low temperatures (Hikosaka et al., 1999, 2006; Sage and Kubien, 2007; Sage et al., 2008). Recently, several researchers indicated that Ar limits the photosynthetic rate at high temperature (Schrader et al., 2004; Wise et al., 2004; Cen and Sage, 2005; Makino and Sage, 2007). They suggested that the deactivation of Rubisco at high temperatures is not the cause of decreased Ac but a result of limitation by Ar. However, it remains unclear whether limitation by Ar is involved in the variation in the optimum temperature for the photosynthetic rate (Fig. 1, B and E).A shift in the optimum temperature for photosynthesis can result from changes in the balance between Ar and Ac, even when the optimum temperatures for these two partial reactions do not change (Fig. 1, C and F; Farquhar and von Caemmerer, 1982). The balance between Ar and Ac has been shown to change depending on growth temperature (Hikosaka et al., 1999; Hikosaka, 2005; Onoda et al., 2005a; Yamori et al., 2005) and often brings about a shift in the colimitation temperature of Ar and Ac. Furthermore, recent studies have shown that plasticity in this balance differs among species or ecotypes (Onoda et al., 2005b; Atkin et al., 2006; Ishikawa et al., 2007). Plasticity in this balance could explain interspecific variation in the plasticity of photosynthetic temperature dependence (Farquhar and von Caemmerer, 1982; Hikosaka et al., 2006), although there has been no evidence in the previous studies that the optimum temperature for photosynthesis occurs at the colimitation point of Ar and Ac.Temperature tolerance differs between species and, with growth temperature, even within species from the same functional group (Long and Woodward, 1989). Bunce (2000) indicated that the temperature dependences of Ar and Ac to growth temperature were different between species from cool and warm climates and that the balance between Ar and Ac was independent of growth temperature for a given plant species. However, it was not clarified what limited the photosynthetic rate or what parameters were important in temperature acclimation of photosynthesis. Recently, we reported that the extent of temperature homeostasis of leaf respiration and photosynthesis, which is assessed as a ratio of rates measured at their respective growth temperatures, differed depending on the extent of the cold tolerance of the species (Yamori et al., 2009b). Therefore, comparisons of several species with different cold tolerances would provide a new insight into interspecific variation of photosynthetic temperature acclimation and their underlying mechanisms. In this study, we selected 11 herbaceous crop species that differ in their cold tolerance (Yamori et al., 2009b) and grew them at two contrasting temperatures, conducting gas-exchange analyses based on the C3 photosynthesis model (Farquhar et al., 1980). Based on these results, we addressed the following key questions. (1) Does the plasticity in photosynthetic temperature acclimation differ between cold-sensitive and cold-tolerant species? (2) Does the limiting step of photosynthesis at several leaf temperatures differ between plant species and with growth temperature? (3) What determines the optimum temperature for the photosynthetic rate among Ac, Ar, and the intersection of the temperature dependences of Ac and Ar?  相似文献   

4.
Projections indicate an elevation of the atmospheric CO2 concentration ([CO2]) concomitant with an intensification of drought for this century, increasing the challenges to food security. On the one hand, drought is a main environmental factor responsible for decreasing crop productivity and grain quality, especially when occurring during the grain-filling stage. On the other hand, elevated [CO2] is predicted to mitigate some of the negative effects of drought. Sorghum (Sorghum bicolor) is a C4 grass that has important economical and nutritional values in many parts of the world. Although the impact of elevated [CO2] and drought in photosynthesis and growth has been well documented for sorghum, the effects of the combination of these two environmental factors on plant metabolism have yet to be determined. To address this question, sorghum plants (cv BRS 330) were grown and monitored at ambient (400 µmol mol−1) or elevated (800 µmol mol−1) [CO2] for 120 d and subjected to drought during the grain-filling stage. Leaf photosynthesis, respiration, and stomatal conductance were measured at 90 and 120 d after planting, and plant organs (leaves, culm, roots, prop roots, and grains) were harvested. Finally, biochemical composition and intracellular metabolites were assessed for each organ. As expected, elevated [CO2] reduced the stomatal conductance, which preserved soil moisture and plant fitness under drought. Interestingly, the whole-plant metabolism was adjusted and protein content in grains was improved by 60% in sorghum grown under elevated [CO2].Global food demand is projected to increase up to 110% by the middle of this century (Tilman et al., 2011; Alexandratos and Bruinsma, 2012), particularly due to a rise in world population that is likely to plateau at about 9 billion people (Godfray et al., 2010). Additionally, the average concentration of atmospheric CO2 ([CO2]) has increased 1.75 µmol mol−1 per year between 1975 and today, reaching 400 µmol mol−1 in April 2015 (NOAA, 2015). According to the A2 emission scenario from the U.S. Environmental Protection Agency, in the absence of explicit climate change policy, atmospheric CO2 concentrations will reach 800 µmol mol−1 by the end of this century. The increasing atmospheric [CO2] is resulting in global climate changes, such as reduction in water availability and elevation in temperature. These factors are expected to heavily influence food production in the next years (Godfray and Garnett, 2014; Magrin et al., 2014).Sorghum (Sorghum bicolor) is a C4 grass, considered a staple food grain for millions of the poorest and most food-insecure people in the semiarid tropics of Africa, Asia, and Central America, serving as an important source of energy, proteins, vitamins, and minerals (Taylor et al., 2006). Moreover, this crop is used for animal feed and as industrial raw material in developed countries such as the United States, which is the main world producer (FAO, 2015). With a fully sequenced genome (Paterson et al., 2009) and over 45,000 accessions representing a large geographic and genetic diversity, sorghum is a good model system in which to study the impact of global climate changes in C4 grasses.The increase in [CO2] in the atmosphere, which is the main driver of global climate changes (Meehl et al., 2007), is predicted to boost photosynthesis rates and productivity in a series of C3 legumes and cereals, mainly due to a decrease in the photorespiration process (Grashoff et al., 1995; Long et al., 2006). On the contrary, due to their capacity to concentrate CO2 in bundle sheath cells and reduce photorespiration to virtually zero, C4 plants are unlikely to respond to the elevation of atmospheric [CO2] (Leakey, 2009). However, even for C4 plants, elevated [CO2] can ameliorate the effects caused by drought, maintaining higher photosynthetic rates. This is due to an improvement in the efficiency of water use that is achieved by the reduction in stomatal conductance (Leakey et al., 2004; Markelz et al., 2011).The rate of photosynthesis as well as the redistribution of photoassimilates accumulated in different plant tissues during the day and/or during vegetative growth are crucial to grain development, and later, to its filling (Schnyder, 1993). Due to this relationship, any environmental stress such as drought occurring during the reproductive phase has the potential to result in poor grain filling and losses in yield (Blum et al., 1997). For instance, postanthesis drought can cause up to 30% decrease in yield (Borrell et al., 2000). It is also known that elevated [CO2], drought, high temperature, and any combinations of these stresses can lead to significant changes in grain composition (Taub et al., 2008; Da Matta et al., 2010; Uprety et al., 2010; Madan et al., 2012), suggesting diverse metabolic alterations and/or adaptations that occur in the plant when it is cultivated in such conditions.Although the impacts of elevated [CO2] and drought on photosynthesis and the growth of sorghum have been well documented (Conley et al., 2001; Ottman et al., 2001; Wall et al., 2001), no attention has been given to the impact of the combination of these two environmental changes on plant metabolism and composition. Regarding physiology, studies on the growth of sorghum under elevated [CO2] and drought showed an increase of the net assimilation rate of 23% due to a decrease of 32% in stomatal conductance (Wall et al., 2001). This resulted in sorghum’s ability to use water 17% more efficiently (Conley et al., 2001). An improvement in the final overall biomass under elevated [CO2] and drought has also been described (Ottman et al., 2001), but without a significant effect in grain yield (Wall et al., 2001).Few studies have been monitoring metabolic pathways in plants under elevated [CO2] (Li et al., 2008; Aranjuelo et al., 2013) and drought (Silvente et al., 2012; Nam et al., 2015; Wenzel et al., 2015). Furthermore, to our knowledge, there are only two reports in which metabolite profiles or metabolic pathways were investigated under the combination of these two environmental conditions (Sicher and Barnaby, 2012; Zinta et al., 2014). Although it is widely accepted that whole-plant metabolism and composition can impact grain filling and yield, metabolic studies conducted so far have focused on a specific plant organ. For instance, Sicher and Barnaby (2012) analyzed the metabolite profile of leaves from maize (Zea mays) plants that were grown under elevated [CO2] and drought, but they did not show how those environmental changes could have affected the metabolism of other tissues (e.g. culm and roots) or how they might have influenced the biomass or grain composition.In order to address how the combination of elevated [CO2] and drought can modify whole-plant metabolism as well as biomass composition in sorghum, this study aimed to (1) evaluate photosynthesis, growth, and yield; (2) underline the differences in biomass composition and primary metabolite profiles among leaves, culm, roots, prop roots, and grains; and (3) determine the effect of elevated [CO2] and drought on the primary metabolism of each organ.  相似文献   

5.
The activation of Rubisco in vivo requires the presence of the regulatory protein Rubisco activase. To elucidate its role in maintaining CO2 assimilation rate at high temperature, we examined the temperature response of CO2 assimilation rate at 380 μL L−1 CO2 concentration (A380) and Rubisco activation state in wild-type and transgenic tobacco (Nicotiana tabacum) with reduced Rubisco activase content grown at either 20°C or 30°C. Analyses of gas exchange and chlorophyll fluorescence showed that in the wild type, A380 was limited by ribulose 1,5-bisphosphate regeneration at lower temperatures, whereas at higher temperatures, A380 was limited by ribulose 1,5-bisphosphate carboxylation irrespective of growth temperatures. Growth temperature induced modest differences in Rubisco activation state that declined with measuring temperature, from mean values of 76% at 15°C to 63% at 40°C in wild-type plants. At measuring temperatures of 25°C and below, an 80% reduction in Rubisco activase content was required before Rubisco activation state was decreased. Above 35°C, Rubisco activation state decreased slightly with more modest decreases in Rubisco activase content, but the extent of the reductions in Rubisco activation state were small, such that a 55% reduction in Rubisco activase content did not alter the temperature sensitivity of Rubisco activation and had no effect on in vivo catalytic turnover rates of Rubisco. There was a strong correlation between Rubisco activase content and Rubisco activation state once Rubisco activase content was less that 20% of wild type at all measuring temperatures. We conclude that reduction in Rubisco activase content does not lead to an increase in the temperature sensitivity of Rubisco activation state in tobacco.The catalytic sites of Rubisco must be activated for CO2 fixation to take place. This requires the carbamylation of a Lys residue at the catalytic sites to allow the binding of Mg2+ and ribulose 1,5-bisphosphate (RuBP; Andrews and Lorimer, 1987). Rubisco activase facilitates carbamylation and the maintenance of Rubisco activity by removing inhibitors such as tight-binding sugar phosphates from Rubisco catalytic sites in an ATP-dependent manner (Andrews, 1996; Spreitzer and Salvucci, 2002; Portis, 2003; Parry et al., 2008). The activity of Rubisco activase is regulated by the ATP/ADP ratio and redox state in the chloroplast (Zhang and Portis, 1999; Zhang et al., 2002; Portis, 2003).In many plant species, Rubisco activation state decreases at high temperature in vivo (Crafts-Brandner and Salvucci, 2000; Salvucci and Crafts-Brandner, 2004b; Cen and Sage, 2005; Yamori et al., 2006b; Makino and Sage, 2007). However, it is unclear what the primary mechanisms underlying the inhibition of Rubisco activation are and whether Rubisco deactivation limits CO2 assimilation rate at high temperature. It has been proposed that Rubisco activation state decreases at high temperature, because the activity of Rubisco activase is insufficient to keep pace with the faster rates of Rubisco inactivation at high temperatures (Crafts-Brandner and Salvucci, 2000; Salvucci and Crafts-Brandner, 2004a, 2004c; Kim and Portis, 2006). In in vitro assays using purified Rubisco and Rubisco activase, the activity of Rubisco activase was sufficient for the activation of Rubisco at the optimum temperature but not at high temperatures (Crafts-Brandner and Salvucci, 2000; Salvucci and Crafts-Brandner, 2004a, 2004c). ATP hydrolysis activity of Rubisco activase in vitro has varying temperature optima among species (e.g. 25°C in Antarctic hairgrass [Deschampsia antarctica] and spinach [Spinacia oleracea] but 35°C in tobacco [Nicotiana tabacum] and cotton [Gossypium hirsutum]), and Rubisco activase more readily dissociates into inactive forms at high temperature, causing a loss of Rubisco activase capacity (Crafts-Brandner and Law, 2000; Salvucci and Crafts-Brandner, 2004b). Moreover, the rates of inhibitor formation by misprotonation of RuBP during catalysis increased at higher temperatures (Salvucci and Crafts-Brandner, 2004c; Kim and Portis, 2006). CO2 assimilation rates and plant growth were improved under heat stress in transgenic Arabidopsis expressing thermotolerant Rubisco activase isoforms generated by either gene-shuffling technology (Kurek et al., 2007) or chimeric Rubisco activase constructs (Kumar et al., 2009). These results support the view that the reduction of Rubisco activase activity limits the Rubisco activation and, therefore, the CO2 assimilation rates at high temperatures.It has also been suggested that the decrease in CO2 assimilation rate at high temperatures is caused by a limitation of RuBP regeneration capacity (e.g. electron transport capacity) rather than by Rubisco deactivation per se (Schrader et al., 2004; Wise et al., 2004; Cen and Sage, 2005; Makino and Sage, 2007; Kubien and Sage, 2008). These groups suggest that Rubisco deactivation at high temperature may be a regulatory response to the limitation of one of the processes contributing to electron transport capacities. For example, at high temperature, protons can leak through the thylakoid membrane, impairing the coupling of ATP synthesis to electron transport (Pastenes and Horton, 1996; Bukhov et al., 1999, 2000). As the electron transport capacity becomes limiting, ATP/ADP ratios and the redox potential of the chloroplast decline, causing a loss of Rubisco activase activity and, in turn, a reduction in the Rubisco activation state (Zhang and Portis, 1999; Zhang et al., 2002; Sage and Kubien, 2007). Based on this understanding, the decline in the Rubisco activation state at high temperature may be a regulated response to a limitation in electron transport capacity rather than a consequence of a direct effect of heat on the integrity of Rubisco activase.Temperature dependence of CO2 assimilation rate shows a considerable variation with growth temperature (Berry and Björkman, 1980; Hikosaka et al., 2006; Sage and Kubien, 2007). Plants grown at low temperature generally exhibit higher CO2 assimilation rates at low temperatures compared with plants grown at high temperature, but they exhibit lower rates at high temperature. Furthermore, both the temperature response of Rubisco activation state and the limiting step of CO2 assimilation rate (a Rubisco versus RuBP regeneration limitation) have been shown to differ depending on growth temperature (Hikosaka et al., 1999; Onoda et al., 2005; Yamori et al., 2005, 2006a, 2006b, 2008). This suggests that the regulation of Rubisco activation state could also differ in plants grown at different growth temperatures. Here, we analyzed the effects of Rubisco activase content on Rubisco activation state and CO2 assimilation rate at leaf temperatures ranging from 15°C to 40°C in tobacco grown under two different temperature regimes (day/night temperatures of 20°C/15°C or 30°C/25°C). We used wild-type and transgenic tobacco with a range of reductions in Rubisco activase content to examine the dependence of Rubisco activation on Rubisco activase content over the range of leaf temperatures (Mate et al., 1993, 1996).  相似文献   

6.
7.
Carbon (C) and nitrogen (N) metabolism are critical to plant growth and development and are at the basis of crop yield and adaptation. We performed high-throughput metabolite analyses on over 12,000 samples from the nested association mapping population to identify genetic variation in C and N metabolism in maize (Zea mays ssp. mays). All samples were grown in the same field and used to identify natural variation controlling the levels of 12 key C and N metabolites, namely chlorophyll a, chlorophyll b, fructose, fumarate, glucose, glutamate, malate, nitrate, starch, sucrose, total amino acids, and total protein, along with the first two principal components derived from them. Our genome-wide association results frequently identified hits with single-gene resolution. In addition to expected genes such as invertases, natural variation was identified in key C4 metabolism genes, including carbonic anhydrases and a malate transporter. Unlike several prior maize studies, extensive pleiotropy was found for C and N metabolites. This integration of field-derived metabolite data with powerful mapping and genomics resources allows for the dissection of key metabolic pathways, providing avenues for future genetic improvement.Carbon (C) and nitrogen (N) metabolism are the basis for life on Earth. The production, balance, and tradeoffs of C and N metabolism are critical to all plant growth, yield, and local adaptation (Coruzzi and Bush, 2001; Coruzzi et al., 2007). In plants, there is a critical balance between the tissues that are producing energy (sources) and those using it (sinks), as the identities and locations of these vary through time and developmental stage (Smith et al., 2004). While a great deal of research has focused on the key genes and proteins involved in these processes (Wang et al., 1993; Kim et al., 2000; Takahashi et al., 2009), relatively little is known about the natural variation within a species that fine-tunes these processes in individual plants.In addition, a key aspect of core C metabolism involves the nature of plant photosynthesis. While the majority of plants use standard C3 photosynthetic pathways, some, including maize (Zea mays) and many other grasses, use C4 photosynthesis to concentrate CO2 in bundle sheath cells to avoid wasteful photorespiration (Sage, 2004). Under some conditions (such as drought or high temperatures), C4 photosynthesis is much more efficient than C3 photosynthesis. Since these conditions are expected to become more prevalent in the near future due to climate change, various research groups are working to convert C3 crop species to C4 metabolism in order to boost crop production and food security (Sage and Zhu, 2011). Beyond this, better understanding of both C3 and C4 metabolic pathways will aid efforts to breed crops for superior yield, N-use efficiency, and other traits important for global food production.In the last two decades, quantitative trait locus (QTL) mapping, first with linkage analysis and later with association mapping, has been used to dissect C and N metabolism in several species, including Arabidopsis (Arabidopsis thaliana; Mitchell-Olds and Pedersen, 1998; Keurentjes et al., 2008; Lisec et al., 2008; Sulpice et al., 2009), tomato (Solanum lycopersicum; Schauer et al., 2006), and maize (Hirel et al., 2001; Limami et al., 2002; Zhang et al., 2006, 2010a, 2010b). These studies identified key genetic regions underlying variation in core C and N metabolism, many of which include candidate genes known to be involved in these processes.Previous studies of genetic variation for C and N metabolism are limited by the fact that they identified trait loci only through linkage mapping in artificial families or through association mapping across populations of unrelated individuals. Linkage mapping benefits from high statistical power due to many individuals sharing the same genotype at any given location, but it suffers from low resolution due to the limited number of generations (and hence recombination events) since the initial founders. Association mapping, in turn, enjoys high resolution due to the long recombination histories of natural populations but suffers from low power, since most genotypes occur in only a few individuals. In addition, many of these studies focused on C and N in artificial settings (e.g. greenhouses or growth chambers) instead of field conditions, running the risk that important genetic loci could be missed if the conditions do not include important (and potentially unknown) natural environmental variables.To address these issues and improve our understanding of C and N metabolism in maize, we used a massive and diverse germplasm resource, the maize nested association mapping (NAM) population (Buckler et al., 2009; McMullen et al., 2009), to evaluate genetic variation underlying the accumulation of 12 targeted metabolites in maize leaf tissue under field conditions. This population was formed by mating 25 diverse maize lines to the reference line, B73, and creating a 200-member biparental family from each of these crosses. The entire 5,000-member NAM population thus combines the strengths of both linkage and association mapping (McMullen et al., 2009), and it has been used to identify QTLs for important traits such as flowering time (Buckler et al., 2009), disease resistance (Kump et al., 2011; Poland et al., 2011), and plant architecture (Tian et al., 2011; Peiffer et al., 2013). Most importantly, this combination of power and resolution frequently resolves associations down to the single-gene level, even when using field-based data.The metabolites we profiled are key indicators of photosynthesis, respiration, glycolysis, and protein and sugar metabolism in the plant (Sulpice et al., 2009). By taking advantage of a robotized metabolic phenotyping platform (Gibon et al., 2004), we performed more than 100,000 assays across 12,000 samples, with two independent samples per experimental plot. Raw data and the best linear unbiased predictors (BLUPs) of these data were included as part of a study of general functional variation in maize (Wallace et al., 2014), but, to our knowledge, this is the first in-depth analysis of these metabolic data. We find strong correlations among several of the metabolites, and we also find extensive pleiotropy among the different traits. Many of the top QTLs are also near or within candidate genes relating to C and N metabolism, thus identifying targets for future breeding and selection. These results provide a powerful resource for those working with core C and N metabolism in plants and for improving maize performance in particular.  相似文献   

8.
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   

9.
10.
The threat to global food security of stagnating yields and population growth makes increasing crop productivity a critical goal over the coming decades. One key target for improving crop productivity and yields is increasing the efficiency of photosynthesis. Central to photosynthesis is Rubisco, which is a critical but often rate-limiting component. Here, we present full Rubisco catalytic properties measured at three temperatures for 75 plants species representing both crops and undomesticated plants from diverse climates. Some newly characterized Rubiscos were naturally “better” compared to crop enzymes and have the potential to improve crop photosynthetic efficiency. The temperature response of the various catalytic parameters was largely consistent across the diverse range of species, though absolute values showed significant variation in Rubisco catalysis, even between closely related species. An analysis of residue differences among the species characterized identified a number of candidate amino acid substitutions that will aid in advancing engineering of improved Rubisco in crop systems. This study provides new insights on the range of Rubisco catalysis and temperature response present in nature, and provides new information to include in models from leaf to canopy and ecosystem scale.In a changing climate and under pressure from a population set to hit nine billion by 2050, global food security will require massive changes to the way food is produced, distributed, and consumed (Ort et al., 2015). To match rising demand, agricultural production must increase by 50 to 70% in the next 35 years, and yet the gains in crop yields initiated by the green revolution are slowing, and in some cases, stagnating (Long and Ort, 2010; Ray et al., 2012). Among a number of areas being pursued to increase crop productivity and food production, improving photosynthetic efficiency is a clear target, offering great promise (Parry et al., 2007; von Caemmerer et al., 2012; Price et al., 2013; Ort et al., 2015). As the gatekeeper of carbon entry into the biosphere and often acting as the rate-limiting step of photosynthesis, Rubisco, the most abundant enzyme on the planet (Ellis, 1979), is an obvious and important target for improving crop photosynthetic efficiency.Rubisco is considered to exhibit comparatively poor catalysis, in terms of catalytic rate, specificity, and CO2 affinity (Tcherkez et al., 2006; Andersson, 2008), leading to the suggestion that even small increases in catalytic efficiency may result in substantial improvements to carbon assimilation across a growing season (Zhu et al., 2004; Parry et al., 2013; Galmés et al., 2014a; Carmo-Silva et al., 2015). If combined with complimentary changes such as optimizing other components of the Calvin Benson or photorespiratory cycles (Raines, 2011; Peterhansel et al., 2013; Simkin et al., 2015), optimized canopy architecture (Drewry et al., 2014), or introducing elements of a carbon concentrating mechanism (Furbank et al., 2009; Lin et al., 2014a; Hanson et al., 2016; Long et al., 2016), Rubisco improvement presents an opportunity to dramatically increase the photosynthetic efficiency of crop plants (McGrath and Long, 2014; Long et al., 2015; Betti et al., 2016). A combination of the available strategies is essential for devising tailored solutions to meet the varied requirements of different crops and the diverse conditions under which they are typically grown around the world.Efforts to engineer an improved Rubisco have not yet produced a “super Rubisco” (Parry et al., 2007; Ort et al., 2015). However, advances in engineering precise changes in model systems continue to provide important developments that are increasing our understanding of Rubisco catalysis (Spreitzer et al., 2005; Whitney et al., 2011a, 2011b; Morita et al., 2014; Wilson et al., 2016), regulation (Andralojc et al., 2012; Carmo-Silva and Salvucci, 2013; Bracher et al., 2015), and biogenesis (Saschenbrecker et al., 2007; Whitney and Sharwood, 2008; Lin et al., 2014b; Hauser et al., 2015; Whitney et al., 2015).A complementary approach is to understand and exploit Rubisco natural diversity. Previous characterization of Rubisco from a limited number of species has not only demonstrated significant differences in the underlying catalytic parameters, but also suggests that further undiscovered diversity exists in nature and that the properties of some of these enzymes could be beneficial if present in crop plants (Carmo-Silva et al., 2015). Recent studies clearly illustrate the variation possible among even closely related species (Galmés et al., 2005, 2014b, 2014c; Kubien et al., 2008; Andralojc et al., 2014; Prins et al., 2016).Until recently, there have been relatively few attempts to characterize the consistency, or lack thereof, of temperature effects on in vitro Rubisco catalysis (Sharwood and Whitney, 2014), and often studies only consider a subset of Rubisco catalytic properties. This type of characterization is particularly important for future engineering efforts, enabling specific temperature effects to be factored into any attempts to modify crops for a future climate. In addition, the ability to coanalyze catalytic properties and DNA or amino acid sequence provides the opportunity to correlate sequence and biochemistry to inform engineering studies (Christin et al., 2008; Kapralov et al., 2011; Rosnow et al., 2015). While the amount of gene sequence information available grows rapidly with improving technology, knowledge of the corresponding biochemical variation resulting has yet to be determined (Cousins et al., 2010; Carmo-Silva et al., 2015; Sharwood and Whitney, 2014; Nunes-Nesi et al., 2016).This study aimed to characterize the catalytic properties of Rubisco from diverse species, comprising a broad range of monocots and dicots from diverse environments. The temperature dependence of Rubisco catalysis was evaluated to tailor Rubisco engineering for crop improvement in specific environments. Catalytic diversity was analyzed alongside the sequence of the Rubisco large subunit gene, rbcL, to identify potential catalytic switches for improving photosynthesis and productivity. In vitro results were compared to the average temperature of the warmest quarter in the regions where each species grows to investigate the role of temperature in modulating Rubisco catalysis.  相似文献   

11.
Selection for cultivars with superior responsiveness to elevated atmospheric CO2 concentrations (eCO2) is a powerful option for boosting crop productivity under future eCO2. However, neither criteria for eCO2 responsiveness nor prescreening methods have been established. The purpose of this study was to identify traits responsible for eCO2 responsiveness of soybean (Glycine max). We grew 12 Japanese and U.S. soybean cultivars that differed in their maturity group and determinacy under ambient CO2 and eCO2 for 2 years in temperature gradient chambers. CO2 elevation significantly increased seed yield per plant, and the magnitude varied widely among the cultivars (from 0% to 62%). The yield increase was best explained by increased aboveground biomass and pod number per plant. These results suggest that the plasticity of pod production under eCO2 results from biomass enhancement, and would therefore be a key factor in the yield response to eCO2, a resource-rich environment. To test this hypothesis, we grew the same cultivars at low planting density, a resource-rich environment that improved the light and nutrient supplies by minimizing competition. Low planting density significantly increased seed yield per plant, and the magnitude ranged from 5% to 105% among the cultivars owing to increased biomass and pod number per plant. The yield increase due to low-density planting was significantly positively correlated with the eCO2 response in both years. These results confirm our hypothesis and suggest that high plasticity of biomass and pod production at a low planting density reveals suitable parameters for breeding to maximize soybean yield under eCO2.The atmospheric concentration of carbon dioxide ([CO2]) increased from the preindustrial level of 271 µmol mol–1 to 391 µmol mol–1 in 2011, owing primarily to emissions from combustion of fossil fuels. [CO2] is predicted to rise from the current level to approximately 600 µmol mol–1 by 2050 (Ciais et al., 2013). Elevated atmospheric CO2 concentration (eCO2) is well known to increase leaf photosynthesis by increasing the availability of CO2 as a substrate for the carboxylation reaction with Rubisco; this can increase crop productivity, a phenomenon known as the CO2 fertilization effect, especially for C3 plants such as rice (Oryza sativa), wheat (Triticum aestivum), and soybean (Glycine max; e.g. Kimball et al., 2002), since [CO2] is a growth-limiting resource for C3 plants. There is a large genotypic variation in the yield response to eCO2, both among cultivars and between species, with responses ranging from –15% to +20% per 100 µmol mol–1 CO2 increase from the current level for rice (Ziska et al., 1996; Moya et al., 1998; Baker, 2004; Shimono et al., 2009; Hasegawa et al., 2013), –6% to +35% for wheat (Manderscheid and Weigel, 1997; Ziska et al., 2004; Ziska, 2008; Tausz-Posch et al., 2015), –5% to +55% for soybean (Ziska and Bunce, 2000; Ziska et al., 2001; Bishop et al., 2015; Bunce, 2015), and –6% to +21% for field bean (Phaseolus vulgaris; Bunce, 2008). These large differences in eCO2 responsiveness within crop species suggest that active selection and breeding for genotypes that respond strongly to gradual but steadily increasing [CO2] may ensure sustained productivity and improve food security in a future eCO2 world (Ainsworth et al., 2008; Ziska et al., 2012; Tausz et al., 2013).Several hypotheses have been proposed about which traits should be targeted by breeders because they are related to intraspecific variation in the responsiveness of seed yield to eCO2. For example, a cultivar’s maturity group is an important growth trait for determining crop productivity. Late-maturing rice cultivars as a result of the longer period in which they can grow could increase grain yield relatively by eCO2 and may therefore benefit more from eCO2 than early-maturing cultivars (Hasegawa et al., 2013). Also, phenological changes by eCO2 could be another good indicator of genotypic variation in eCO2 responsiveness. Recently, Bunce (2015) showed that extension of the duration of vegetative growth until flowering at the apical node of the main stem caused by eCO2 was correlated with an increase in seed yield among some soybean genotypes.The source-sink relationship is another important aspect of genotypic variation in the responsiveness of a plant to eCO2. CO2 enrichment can increase photosynthesis, especially during the early growth stage of leaves, and the magnitude of the increase of photosynthesis decreases with increasing growth stage because leaf senescence accelerates under eCO2, which has been referred to as acclimation (for review, see Moore et al., 1999). Genotypes with slower acclimation to eCO2 had a greater response of seed yield (Zhu et al., 2014, for rice; Hao et al., 2012, for soybean). Determinacy is strongly related to the source-sink relationship, especially for legume species. Indeterminate soybean cultivars that have more sinks are likely to be superior in responsiveness to eCO2, compared with determinate cultivars (Ainsworth et al., 2004). Aspinwall et al. (2015) emphasized the importance of phenotypic plasticity (the ability of a genotype to alter its phenotype in response to environmental changes) under eCO2 as a key trait for eCO2 responsiveness. Many researchers have suggested that a higher sink plasticity under eCO2 would lead to greater plasticity of tillering, branching, and biomass production, and could therefore be more important than photosynthesis per unit leaf area for adaptation to eCO2 by rice (Shimono et al., 2009; Zhu et al., 2014), soybean (Ziska and Bunce, 2000; Ziska et al., 2001), wheat (Manderscheid and Weigel, 1997; Ziska et al., 2004; Ziska, 2008), and field bean (Bunce, 2008); an exception would occur when other resources are limited, such as during a drought (Tausz-Posch et al., 2015). It is difficult to categorize biomass per se as a function of sink or source factors, but a plant’s final biomass results from efficient formation of sinks in vegetative and reproductive organs, and from efficient filling of these sinks with the products of photosynthesis. This would be a promising hypothesis, and if the hypothesis is confirmed, the phenotypic plasticity would become a useful criterion for identifying eCO2-responsive cultivars.The first objective of the current study was to examine the genotypic variation of yield enhancement caused by eCO2 in diverse soybean cultivars, as soybean is a major source of plant protein and oil and a major contributor to the world’s food supply. In addition to characterizing the variation, we attempted to identify the factors responsible for it. The soybean genotypes that we chose covered a wide range of maturity groups and determinacy, including near-isogenic lines. We concluded that soybean cultivars varied widely in their responsiveness to eCO2, and that variation in the yield enhancement by eCO2 was determined primarily by the plasticity of biomass and pod production. This suggests that both are suitable parameters for screening cultivars with a strong response to eCO2. However, it is not easy to characterize plasticity of biomass and pod production under eCO2 of each cultivar since CO2 enhancement facilities such as controlled enclosed chambers are extremely expensive and not easily accessible.Our second objective was to develop a methodology to characterize intraspecific variation in plasticity. Shimono (2011) proposed a simple and novel idea: to use planting density for prescreening to identify eCO2-responsive cultivars. Solar radiation is the driving force for photosynthesis and plant growth, and crops are grown as a population (not as individual plants) to maximize productivity per unit area rather than per plant. Individual plant growth is usually restricted by interplant competition for solar radiation, soil nutrients, and water, so a lower planting density could potentially increase the source strength for individual plants by increasing the availability of resources. Thus, lower plant density can imitate the greater resource availability that would occur under eCO2. This approach is potentially useful but is insufficient, because Shimono et al. (2014) applied it only to rice, measured only panicle number (not yield) as the phenotype, and combined independent experimental data from different locations and years. Here, we tested the hypothesis using a diverse range of soybean cultivars at a single site and for 2 years, and found a good relationship between the responsiveness to eCO2 and the responsiveness to low planting density (LD) in terms of the seed yield per plant.  相似文献   

12.
13.
14.
15.
Rising global temperature and CO2 levels may sustain late-season net photosynthesis of evergreen conifers but could also impair the development of cold hardiness. Our study investigated how elevated temperature, and the combination of elevated temperature with elevated CO2, affected photosynthetic rates, leaf carbohydrates, freezing tolerance, and proteins involved in photosynthesis and cold hardening in Eastern white pine (Pinus strobus). We designed an experiment where control seedlings were acclimated to long photoperiod (day/night 14/10 h), warm temperature (22°C/15°C), and either ambient (400 μL L−1) or elevated (800 μmol mol−1) CO2, and then shifted seedlings to growth conditions with short photoperiod (8/16 h) and low temperature/ambient CO2 (LTAC), elevated temperature/ambient CO2 (ETAC), or elevated temperature/elevated CO2 (ETEC). Exposure to LTAC induced down-regulation of photosynthesis, development of sustained nonphotochemical quenching, accumulation of soluble carbohydrates, expression of a 16-kD dehydrin absent under long photoperiod, and increased freezing tolerance. In ETAC seedlings, photosynthesis was not down-regulated, while accumulation of soluble carbohydrates, dehydrin expression, and freezing tolerance were impaired. ETEC seedlings revealed increased photosynthesis and improved water use efficiency but impaired dehydrin expression and freezing tolerance similar to ETAC seedlings. Sixteen-kilodalton dehydrin expression strongly correlated with increases in freezing tolerance, suggesting its involvement in the development of cold hardiness in P. strobus. Our findings suggest that exposure to elevated temperature and CO2 during autumn can delay down-regulation of photosynthesis and stimulate late-season net photosynthesis in P. strobus seedlings. However, this comes at the cost of impaired freezing tolerance. Elevated temperature and CO2 also impaired freezing tolerance. However, unless the frequency and timing of extreme low-temperature events changes, this is unlikely to increase risk of freezing damage in P. strobus seedlings.Land surface temperature is increasing, particularly in the northern hemisphere (IPCC, 2014), which is dominated by boreal and temperate forests. At higher latitudes, trees rely on temperature and photoperiod cues to detect changing seasons and to trigger cessation of growth and cold hardening during the autumn (Ensminger et al., 2015). For boreal and temperate evergreen conifers, cold hardening involves changes in carbohydrate metabolism, down-regulation of photosynthesis, accumulation of cryoprotective metabolites, and development of freezing tolerance (Crosatti et al., 2013; Ensminger et al., 2015). These processes minimize freezing damage and enable conifers to endure winter stresses. However, rising temperatures result in asynchronous phasing of temperature and photoperiod characterized by delayed arrival of first frosts (McMahon et al., 2010), which may impact the onset and development of cold hardening during autumn.Short photoperiod induces the cessation of growth in many tree species (Downs and Borthwick, 1956; Heide, 1974; Repo et al., 2000; Böhlenius et al., 2006). As a consequence, carbon demand in sink tissue decreases toward the end of the growing season, and the bulk of photoassimilate is translocated from source tissues to storage tissues (Hansen and Beck, 1994; Oleksyn et al., 2000). In addition, cryoprotective soluble sugars, including sucrose, raffinose, and pinitol, accumulate in leaf tissues to enhance freezing tolerance (Strimbeck et al., 2008; Angelcheva et al., 2014). Thus, by winter, leaf nonstructural carbohydrates are mainly comprised of mono- and oligosaccharides, and only minimal levels of starch remain (Hansen and Beck, 1994; Strimbeck et al., 2008). The concurrent decrease of photoassimilate and demand for metabolites that occur during the cessation of growth also impacts the citric acid cycle that mediates between photosynthesis, respiration, and protein synthesis. The citric acid cycle generates NADH to fuel ATP synthesis via mitochondrial electron transport, as well as amino acid precursors (Shi et al., 2015). In C3 plants, the enzyme phosphoenolpyruvate carboxylase (PEPC) converts phosphoenolpyruvate to oxaloacetic acid in order to supplement the flow of metabolites to the citric acid cycle and thus controls the regulation of respiration and photosynthate partitioning (O’Leary et al., 2011).Cessation of growth, low temperature, and presumably short photoperiod decrease the metabolic sink for photoassimilates, resulting in harmful excess light energy (Öquist and Huner, 2003; Ensminger et al., 2006) and increased generation of reactive oxygen species (Adams et al., 2004). During autumn and the development of cold hardiness, conifers reconfigure the photosynthetic apparatus in order to avoid formation of excess light and reactive oxygen species. This involves a decrease in chlorophylls and PSII reaction center core protein D1 (Ottander et al., 1995; Ensminger et al., 2004; Verhoeven et al., 2009), as well as aggregation of light-harvesting complex proteins (Ottander et al., 1995; Busch et al., 2007). Additionally, photoprotective carotenoid pigments accumulate in leaves, especially the xanthophylls, zeaxanthin, and lutein that contribute to nonphotochemical quenching (NPQ) via thermal dissipation of excess light energy (Busch et al., 2007; Verhoeven et al., 2009; Demmig-Adams et al., 2012). Prolonged exposure to low temperature induces sustained nonphotochemical quenching (NPQS), where zeaxanthin constitutively dissipates excess light energy (Ensminger et al., 2004; Demmig-Adams et al., 2012; Fréchette et al., 2015).In conifers, freezing tolerance is initiated during early autumn in response to decreasing photoperiod (Rostad et al., 2006; Chang et al., 2015) and continues to develop through late autumn in response to the combination of short photoperiod and low temperature (Strimbeck and Schaberg, 2009; Chang et al., 2015). In addition to changes in carbohydrate content, freezing tolerance also involves the expression of specific dehydrins (Close, 1997; Kjellsen et al., 2013). Members of the dehydrin protein family are involved in responses to osmotic, salt, and freezing stress (Close, 1996). Dehydrins have been associated with improved freezing tolerance in many species including spinach (Kaye et al., 1998), strawberry (Houde et al., 2004), cucumber (Yin et al., 2006), peach (Wisniewski et al., 1999), birch (Puhakainen et al., 2004), and spruce (Kjellsen et al., 2013). In angiosperms, a characteristic Lys-rich dehydrin motif known as the K-segment interacts with lipids to facilitate membrane binding (Koag et al., 2003; Eriksson et al., 2011). Several in vitro studies have demonstrated dehydrin functions including prevention of aggregation and unfolding of enzymes (using Vitis riparia; Hughes and Graether, 2011), radical scavenging (using Citrus unshiu; Hara et al., 2004), and suppression of ice crystal formation (using Prunus persica; Wisniewski et al., 1999). To date, dehydrin functions have not been demonstrated in planta.Rising temperatures since the mid-twentieth century have delayed the onset of autumn dormancy and increased length of the growing season in forests across the northern hemisphere (Boisvenue and Running, 2006; Piao et al., 2007; McMahon et al., 2010). Studies have shown that elevated temperatures ranging from +4°C to +20°C above ambient can delay down-regulation of photosynthesis in several evergreen conifers. Consistent findings were apparent among climate-controlled chamber studies exposing Pinus strobus seedlings to a sudden shift in temperature and/or photoperiod (Fréchette et al., 2016), as well as chamber studies exposing Picea abies seedlings to simulated autumn conditions using a gradient of decreasing temperature and photoperiod (Stinziano et al., 2015). Similar findings were also demonstrated in open-top chamber experiments exposing mature Pinus sylvestris to a gradient of decreasing temperature and natural photoperiod (Wang, 1996). Elevated temperature (+4°C above ambient) also impaired cold hardening in Pseudotsuga menziesii seedlings (Guak et al., 1998) and mature P. sylvestris (Repo et al., 1996) exposed to a decreasing gradient of temperature and natural photoperiod using open-top chambers. In contrast, a recent study showed that smaller temperature increments (+1.5°C to +3°C) applied using infrared heaters did not delay down-regulation of photosynthesis or impair freezing tolerance in field-grown P. strobus seedlings that were acclimated to larger diurnal and seasonal temperature variations (Chang et al., 2015). For many tree species, photoperiod determines cessation of growth (Tanino et al., 2010; Petterle et al., 2013), length of the growing season (Bauerle et al., 2012), and development of cold hardiness (Welling et al., 1997; Li et al., 2003; Rostad et al., 2006). However, the effects of climate warming on tree phenology are complex and can be unpredictable due to species- and provenance-specific differences in sensitivity to photoperiod and temperature cues (Körner and Basler, 2010; Basler and Körner, 2012; Basler and Körner, 2014).The effect of elevated CO2 further increases uncertainties in the response of trees to warmer climate. Similar to warmer temperature, elevated CO2 may also delay the down-regulation of photosynthesis in evergreens and extend the length of the growing season, as demonstrated in mature P. sylvestris (Wang, 1996). Elevated CO2 increases carbon assimilation (Curtis and Wang, 1998; Ainsworth and Long, 2005) and biomass production (Ainsworth and Long, 2005) during the growing season. The effects could continue during the autumn if dormancy or growth cessation is delayed, which suggests that elevated CO2 may increase annual carbon uptake. However, long-term exposure to elevated CO2 can also down-regulate photosynthesis during the growing season (Ainsworth and Long, 2005). Prior studies that have attempted to determine the impact of a combination of elevated CO2 and/or temperature on cold hardening in evergreens have largely focused on freezing tolerance, with contrasting results. Open-top chamber experiments showed that a combination of elevated temperature and CO2 both delayed and impaired freezing tolerance of P. menziesii seedlings (Guak et al., 1998) and evergreen broadleaf Eucalyptus pauciflora seedlings (Loveys et al., 2006) but did not affect freezing tolerance of mature P. sylvestris (Repo et al., 1996). A recent field experiment examining mature trees revealed that Larix decidua, but not Pinus mugo, exhibited enhanced freezing damage following six years of exposure to combined soil warming and elevated CO2 (Rixen et al., 2012). In contrast, a climate-controlled study showed that exposure to elevated CO2 advanced the date of bud set and improved freezing tolerance in Picea mariana seedlings (Bigras and Bertrand, 2006). In a second study on similar seedlings conducted by the same authors, exposure of trees to elevated CO2 also enhanced freezing tolerance but impaired the accumulation of sucrose and raffinose (Bertrand and Bigras, 2006). These previous experiments used experimental conditions where temperature and photoperiod gradually decreased. While this approach aims to mimic natural conditions, it is difficult to distinguish specific responses to either photoperiod or temperature. Because of the contrasting findings from previous studies, we designed an experiment aiming to separate the effects of photoperiod, temperature, and CO2 on a wide range of parameters that are involved in cold hardening in conifers.Our study aimed to determine (1) how induction and development of the cold hardening process is affected by a shift from long to short photoperiod under warm conditions and (2) how the combination of warm air temperature and elevated CO2 affects photoperiod-induced cold hardening processes in Eastern white pine (P. strobus). To assess the development of cold hardening, we measured photosynthetic rates, changes in leaf carbohydrates, freezing tolerance, and proteins involved in photosynthesis and cold hardening over 36 d. Assuming that both low temperature and short photoperiod cues are required to induce cold hardening in conifers, we hypothesized that warm temperature and the combination of warm temperature and elevated CO2 would prevent seedlings growing under autumn photoperiod from down-regulating photosynthesis. We further hypothesized that warm temperature and the combination of warm temperature and elevated CO2 would impair the development of freezing tolerance, due to a lack of adequate phasing of the low temperature and short photoperiod signals.  相似文献   

16.
The conversion efficiency (εc) of absorbed radiation into biomass (MJ of dry matter per MJ of absorbed photosynthetically active radiation) is a component of yield potential that has been estimated at less than half the theoretical maximum. Various strategies have been proposed to improve εc, but a statistical analysis to establish baseline εc levels across different crop functional types is lacking. Data from 164 published εc studies conducted in relatively unstressed growth conditions were used to determine the means, greatest contributors to variation, and genetic trends in εc across important food and biofuel crop species. εc was greatest in biofuel crops (0.049–0.066), followed by C4 food crops (0.046–0.049), C3 nonlegumes (0.036–0.041), and finally C3 legumes (0.028–0.035). Despite confining our analysis to relatively unstressed growth conditions, total incident solar radiation and average growing season temperature most often accounted for the largest portion of εc variability. Genetic improvements in εc, when present, were less than 0.7% per year, revealing the unrealized potential of improving εc as a promising contributing strategy to meet projected future agricultural demand.Substantial increases in yield are needed to feed and fuel the world’s growing human population. With an estimated population of nine billion people by the middle of this century (Lutz and Samir, 2010) and rising affluence resulting in greater consumption of grain-fed animal products (Cirera and Masset, 2010), different studies predict that, by midcentury, global crop production will need to increase 60% to 120% over 2005 levels without the expansion of agricultural land area (Tilman et al., 2011; Alexandratos and Bruinsma, 2012).Doubling yields in major food and fuel crops requires considerable effort, especially as yields are beginning to plateau in many major food crops. Yield increases necessary for doubling productivity by midcentury are estimated at 1.16% to 1.31% each year in all cereals (Hall and Richards, 2013), 1.7% per year in wheat (Triticum aestivum; Rosegrant and Agcaoili, 2010), and 2.4% (noncompounding average per year) across all major grain crops (Ray et al., 2013). However, global mean increases from the past 20 to 30 years suggest that yield gains in rice (Oryza sativa) and wheat are approximately 1% (Lopes et al., 2012; Manès et al, 2012; Ray et al., 2013) and declining in some areas of the world (Cassman et al., 2010; Fischer and Edmeades, 2010; Long and Ort, 2010; Ray et al., 2013). Global yearly increases are estimated at 1.3% in soybean (Glycine max) and 1.6% in maize (Zea mays), with similar concerns that yield trends may also be decreasing in some major growing regions (Lobell and Gourdji, 2012; Ray et al., 2013).Efforts to increase yields in the next few decades must also account for environmental and sustainability goals (Sayer et al., 2013) as well as heightened environmental stresses predicted to occur due to climate change, which are already responsible for some of the stagnation in yield increases. Anthropogenic sources of greenhouse gases have caused an approximately 1°C increase in land surface temperatures since 1900, and global mean surface temperatures are likely to increase by up to 2.4°C to 4.8°C by the end of the century (IPCC, 2013). Drought is also expected to become more frequent and intense in many regions of the world (Dai, 2011; IPCC, 2013). Of the variability present in major food crop yield gains, 30% can be explained by climate change alone (Lobell and Field, 2007), with drastic decreases in barley (Hordeum vulgare), maize, rice, sorghum (Sorghum bicolor), soybean, and wheat yields as average growing season temperatures surpass the temperature optimum for each crop (Lobell and Gourdji, 2012). Current levels of atmospheric CO2 concentration [CO2] are the highest they have been in at least 800,000 years (IPCC, 2013). Elevated [CO2] increases water use efficiency (Ainsworth and Long, 2005, Bernacchi et al., 2007, Leakey et al., 2009), but probably not to an extent that would mitigate the resulting reductions in yield caused by higher temperature and higher vapor pressure deficit (Ort and Long, 2014). Additionally, any fertilization effects on C3 yields due to elevated [CO2] would be at least in part negated by drought and temperature stress, leaving yield increases far from optimal (Long et al., 2006a; Lobell and Gourdji, 2012).  相似文献   

17.
Unequal absorption of photons between photosystems I and II, and between bundle-sheath and mesophyll cells, are likely to affect the efficiency of the CO2-concentrating mechanism in C4 plants. Under steady-state conditions, it is expected that the biochemical distribution of energy (ATP and NADPH) and photosynthetic metabolite concentrations will adjust to maintain the efficiency of C4 photosynthesis through the coordination of the C3 (Calvin-Benson-Bassham) and C4 (CO2 pump) cycles. However, under transient conditions, changes in light quality will likely alter the coordination of the C3 and C4 cycles, influencing rates of CO2 assimilation and decreasing the efficiency of the CO2-concentrating mechanism. To test these hypotheses, we measured leaf gas exchange, leaf discrimination, chlorophyll fluorescence, electrochromatic shift, photosynthetic metabolite pools, and chloroplast movement in maize (Zea mays) and Miscanthus × giganteus following transitional changes in light quality. In both species, the rate of net CO2 assimilation responded quickly to changes in light treatments, with lower rates of net CO2 assimilation under blue light compared with red, green, and blue light, red light, and green light. Under steady state, the efficiency of CO2-concentrating mechanisms was similar; however, transient changes affected the coordination of C3 and C4 cycles in M. giganteus but to a lesser extent in maize. The species differences in the ability to coordinate the activities of C3 and C4 cycles appear to be related to differences in the response of cyclic electron flux around photosystem I and potentially chloroplast rearrangement in response to changes in light quality.The CO2-concentrating mechanism in C4 plants reduces the carbon lost through the photorespiratory pathway by limiting the oxygenation of ribulose-1,5-bisphosphate (RuBP) by the enzyme Rubisco (Brown and Smith, 1972; Sage, 1999). Through the compartmentalization of the C4 cycle in the mesophyll cells and the C3 cycle in the bundle-sheath cells (Hatch and Slack, 1966), C4 plants suppress RuBP oxygenation by generating a high CO2 partial pressure around Rubisco (Furbank and Hatch, 1987). To maintain high photosynthetic rates and efficient light energy utilization, the metabolic flux through the C3 and C4 cycles must be coordinated. However, coordination of the C3 and C4 cycles is likely disrupted due to rapid changes in environmental conditions, particularly changes in light availability (Evans et al., 2007; Tazoe et al., 2008).Spatial and temporal variations in light environments, including both light quantity and quality, are expected to alter the coordination of the C3 and C4 cycles. For example, it has been suggested that the coordination of C3 and C4 cycles is altered by changes in light intensity (Henderson et al., 1992; Cousins et al., 2006; Tazoe et al., 2006, 2008; Kromdijk et al., 2008, 2010; Pengelly et al., 2010). However, more recent publications indicate that some of the proposed light sensitivity of the CO2-concentrating mechanisms in C4 plants can be attributed to oversimplifications of leaf models of carbon isotope discrimination (Δ13C), in particular, errors in estimates of bundle-sheath CO2 partial pressure and omissions of respiratory fractionation (Ubierna et al., 2011, 2013). Alternatively, there is little information on the effects of light quality on the coordination of C3 and C4 cycle activities and the subsequent impact on net rate of CO2 assimilation (Anet).In C3 plants, Anet is reduced under blue light compared with red or green light (Evans and Vogelmann, 2003; Loreto et al., 2009). This was attributed to differences in absorbance and wavelength-dependent differences in light penetration into leaves, where red and green light penetrate farther into leaves compared with blue light (Vogelmann and Evans, 2002; Evans and Vogelmann, 2003). Differences in light quality penetration into a leaf are likely to have profound impacts on C4 photosynthesis, because the C4 photosynthetic pathway requires the metabolic coordination of the mesophyll C4 cycle and the bundle-sheath C3 cycle. Indeed, Evans et al. (2007) observed a 50% reduction in the rate of CO2 assimilation in Flaveria bidentis under blue light relative to white light at a light intensity of 350 µmol quanta m−2 s−1. This was attributed to poor penetration of blue light into the bundle-sheath cells and subsequent insufficient production of ATP in the bundle-sheath cells to match the rates of mesophyll cell CO2 pumping (Evans et al., 2007). Recently, Sun et al. (2012) observed similar low rates of steady-state CO2 assimilation under blue light relative to red, green, and blue light (RGB), red light, and green light at a constant light intensity of 900 µmol quanta m−2 s−1.Because the light penetration into a leaf depends on light quality, with blue light penetrating the least, this potentially results in changes in the energy available for carboxylation reactions in the bundle-sheath (C3 cycle) and mesophyll (C4 cycle) cells. Changes in the balance of energy driving the C3 and C4 cycles can alter the efficiency of the CO2-concentrating mechanisms, often represented by leakiness (ϕ), the fraction of CO2 that is pumped into the bundle-sheath cells that subsequently leaks back out (Evans et al., 2007). Unfortunately, ϕ cannot be measured directly, but it can be estimated through the combined measured and modeled values of Δ13C (Farquhar, 1983). Using measurements of Δ13C, it has been demonstrated that under steady-state conditions, changes in light quality do not affect ϕ (Sun et al., 2012); however, it remains unknown if ϕ is also constant during the transitions between different light qualities. In fact, sudden changes of light quality could temporally alter the coordination of the C3 and C4 cycles.To understand the effects of light quality on C4 photosynthesis and the coordination of the activities of C3 and C4 cycles, we measured transitional changes in leaf gas exchange and Δ13C under RGB and broad-spectrum red, green, and blue light in the NADP-malic enzyme C4 plants maize (Zea mays) and Miscanthus × giganteus. Leaf gas exchange and Δ13C measurements were used to estimate ϕ using the complete model of C4 leaf Δ13C (Farquhar, 1983; Farquhar and Cernusak, 2012). Additionally, we measured photosynthetic metabolite pools, Rubisco activation state, chloroplast movement, and rates of linear versus cyclic electron flow during rapid transitions from red to blue light and blue to red light. We hypothesized that the limited penetration of blue light into the leaf would result in insufficient production of ATP in the bundle-sheath cells to match the rate of mesophyll cell CO2 pumping. We predicted that rapid changes in light quality would affect the coordination of the C3 and C4 cycles and cause an increase in ϕ, but this would equilibrate as leaf metabolism reached a new steady-state condition.  相似文献   

18.
An antisense construct targeting the C4 isoform of NADP-malic enzyme (ME), the primary enzyme decarboxylating malate in bundle sheath cells to supply CO2 to Rubisco, was used to transform the dicot Flaveria bidentis. Transgenic plants (α-NADP-ME) exhibited a 34% to 75% reduction in NADP-ME activity relative to the wild type with no visible growth phenotype. We characterized the effect of reducing NADP-ME on photosynthesis by measuring in vitro photosynthetic enzyme activity, gas exchange, and real-time carbon isotope discrimination (Δ). In α-NADP-ME plants with less than 40% of wild-type NADP-ME activity, CO2 assimilation rates at high intercellular CO2 were significantly reduced, whereas the in vitro activities of both phosphoenolpyruvate carboxylase and Rubisco were increased. Δ measured concurrently with gas exchange in these plants showed a lower Δ and thus a lower calculated leakiness of CO2 (the ratio of CO2 leak rate from the bundle sheath to the rate of CO2 supply). Comparative measurements on antisense Rubisco small subunit F. bidentis plants showed the opposite effect of increased Δ and leakiness. We use these measurements to estimate the C4 cycle rate, bundle sheath leak rate, and bundle sheath CO2 concentration. The comparison of α-NADP-ME and antisense Rubisco small subunit demonstrates that the coordination of the C3 and C4 cycles that exist during environmental perturbations by light and CO2 can be disrupted through transgenic manipulations. Furthermore, our results suggest that the efficiency of the C4 pathway could potentially be improved through a reduction in C4 cycle activity or increased C3 cycle activity.In the leaves of a range of plants including maize (Zea mays), sorghum (Sorghum bicolor), sugarcane (Saccharum officinarum), and millet (Pennisetum americanum), a biochemical pathway known as C4 photosynthesis has evolved to concentrate CO2 at the site of Rubisco such that Rubisco can operate at close to its maximal activity and photorespiration is reduced, enhancing the rate of photosynthesis in air (Hatch, 1987; Sage, 2004). In most C4 plants, CO2 is fixed by phosphoenolpyruvate carboxylase (PEPC) in the mesophyll cells into four-carbon acids, which diffuse to an inner ring of bundle sheath cells, where they are decarboxylated and the CO2 is refixed by Rubisco. Plants using the C4 photosynthetic mechanism have been subdivided into three primary subtypes, the NADP-malic enzyme (ME), NAD-ME, and phosphoenolpyruvate carboxykinase types, according to the decarboxylating enzyme used to generate CO2 from C4 acids in the bundle sheath cells (Hatch, 1987). Flaveria bidentis is a typical NADP-ME dicot in which malate and Asp contribute equally in the transfer of CO2 to bundle sheath cells (Meister et al., 1996). Presumably, in most C4 plants, the reactions that facilitate the appropriation, transformation, transport, and eventual concentration of CO2 in the bundle sheath cell chloroplasts (C4 cycle) are balanced with the reactions that incorporate CO2 into usable carbon compounds for energy (C3/Calvin cycle) such that energy is not lost or wasted as environmental conditions fluctuate. This process is important in maintaining the efficiency of the CO2-concentrating mechanism and of C4 photosynthesis overall. The nature of the controlling mechanisms for balance and coordination between the C3 and C4 cycles is still unclear, however, and concrete evidence for the coordinated regulation of primary carboxylation in the mesophyll and decarboxylation of C4 acids in the bundle sheath has not been forthcoming. A key approach to revealing these mechanisms has been the use of antisense RNA in the C4 dicot F. bidentis to reduce levels of key photosynthetic enzymes, including Rubisco (Furbank et al., 1996), NADP-malate dehydrogenase and pyruvate phosphate dikinase (Furbank et al., 1997), Rubisco activase (von Caemmerer et al., 2005), carbonic anhydrase (Cousins et al., 2006), and PEPC protein kinase (Furumoto et al., 2007). This has proven to be a valuable method to help gain insight into enzyme function and regulation during C4 photosynthesis and to potentially alter the balance between the C3 and C4 cycles.In this study, we targeted the gene encoding the chloroplastic C4 isozyme of NADP-ME in F. bidentis (Marshall et al., 1996) with an antisense construct designed to reduce its activity in vivo. This isoform is thought to catalyze the decarboxylation of l-malate to pyruvate and CO2 and of NADP to NADPH in bundle sheath chloroplasts during C4 photosynthesis (Ashton, 1997; Drincovich et al., 2001), allowing the CO2 to be fixed into the C3 cycle by Rubisco and pyruvate to return back to mesophyll cells to be recycled into PEP. These antisense lines were generated for two purposes. First, these plants could be used to confirm the identity of the gene encoding the NADP-ME isozyme involved in C4 photosynthesis. Several other functioning isoforms of NADP-ME have also been identified within Flaveria spp.: a chloroplastic but potentially nonphotosynthetic NADP-ME form and a cytosolic NADP-ME (Marshall et al., 1996; Drincovich et al., 1998; Lai et al., 2002). The specific role and regulation of a C4 NADP-ME isozyme in F. bidentis is of interest in relation to the “transfer” or “generation” of a functioning C4 cycle to C3 plants (Sheehy et al., 2007; Furbank et al., 2009). A greater understanding of the balance and interactions between this enzyme and others in the C4 and C3 cycles will aid in deciding the expression locations and levels needed for C3 plants to gain a functional CO2-concentrating mechanism.The second use of these antisense plants was to investigate the degree of coordination between the C4/C3 cycles in F. bidentis and the possibility of manipulation to improve photosynthetic efficiency. As mentioned above, the mechanisms of regulation (if any) of the C3 pathway enzymes such as Rubisco in response to the activity and CO2 supply rate of the C4 cycle are unknown. It is similarly unclear how much the reactions of the C3 cycle affect the rates of the initial CO2-fixing reactions (carbonic anhydrase and PEPC). Leakiness (ϕ), defined as the ratio of CO2 leak rate from the bundle sheath to the rate of CO2 supply, reflects the coordination of the C4 and C3 cycles by describing the amount of overcycling of the C4 cycle that has to occur to support a given rate of net CO2 assimilation (Furbank et al., 1990; von Caemmerer and Furbank, 1999). As a major C4 enzyme functioning within the bundle sheath, a reduction in NADP-ME should affect both the C4 cycle rate and the bundle sheath CO2 concentration (Cs), possibly disrupting the enzymatic balance and coordination in F. bidentis. Here, we have designed experiments to simultaneously look at in vitro photosynthetic enzyme activity, gas exchange, and real-time carbon isotope discrimination (Δ), facilitating estimates of ϕ, C4 cycle rate, and the possible range of Cs within transgenic α-NADP-ME and antisense Rubisco small subunit (α-SSu) F. bidentis plants (Furbank et al., 1996). These measurements aim to show the impact of our perturbations of the C3/C4 balance, highlighting possible communication pathways between the cycles and also other possible targets for future genetic manipulation to improve the rate and/or efficiency of photosynthesis in C4 plants.  相似文献   

19.
20.
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   

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