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 328 毫秒
1.
Vertical leaf nitrogen (N) gradient within a canopy is classically considered as a key adaptation to the local light environment that would tend to maximize canopy photosynthesis. We studied the vertical leaf N gradient with respect to the light gradient for wheat (Triticum aestivum) canopies with the aims of quantifying its modulation by crop N status and genetic variability and analyzing its ecophysiological determinants. The vertical distribution of leaf N and light was analyzed at anthesis for 16 cultivars grown in the field in two consecutive seasons under two levels of N. The N extinction coefficient with respect to light (b) varied with N supply and cultivar. Interestingly, a scaling relationship was observed between b and the size of the canopy for all the cultivars in the different environmental conditions. The scaling coefficient of the b-green area index relationship differed among cultivars, suggesting that cultivars could be more or less adapted to low-productivity environments. We conclude that the acclimation of the leaf N gradient to the light gradient is a whole-plant process that depends on canopy size. This study demonstrates that modeling leaf N distribution and canopy expansion based on the assumption that leaf N distribution parallels that of the light is inappropriate. We provide a robust relationship accounting for vertical leaf N gradient with respect to vertical light gradient as a function of canopy size.In cereals, as in many crop species, nitrogen (N) nutrition is a major determinant in the elaboration of grain yield and quality (Lemaire and Millard, 1999; Lawlor, 2002; Hikosaka, 2005). N is involved in both meristematic and photosynthetic activities, with consequences on plant architecture and carbon acquisition and in fine on grain yield and protein concentration. Beside the total amount of N absorbed by the crop, the allocation of N among plant organs plays a key role in determining crop productivity and quality (Grindlay, 1997; Dreccer et al., 1998; Hikosaka, 2005).Light interception and leaf N content are the two main factors governing carbon assimilation at the leaf scale (Evans, 1989). For various species, both light and leaf N attenuate with cumulative leaf area index counted from the top of the canopy (Field, 1983; Hirose and Werger, 1987). Leaf N vertical gradients have been regarded as an adaptive response to the local light environment, maximizing canopy photosynthesis and N utilization efficiency (Hirose and Werger, 1987; Hikosaka et al., 1994; Drouet and Bonhomme, 1999), as N is largely contained in the assimilatory enzyme Rubisco. Theoretical studies indicated that leaf N maximizes canopy photosynthesis when it parallels the light gradient (i.e. when the light [KL] and N [KN] extinction coefficients are equal), considering that the leaf N gradient is “optimal” in accordance with the “optimization theory” (Field, 1983; Hirose and Werger, 1987; Anten et al., 1995b).Factors other than the photosynthetic photon flux density (PPFD) might be responsible for the observed leaf N distribution. For instance, the acropetal gradients of leaf age (Hikosaka et al., 1994; Hikosaka, 2005) and light composition (Rousseaux et al., 1999) are known to strengthen the leaf N gradient. However, the impact of each of these factors has been shown to be much less than that of the PPFD gradient (Werger and Hirose, 1991; Pons and de Jong-van Berkel, 2004), although for the grass species Brachypodium pinnatum other factors than light might be involved (Pons et al., 1993). At the molecular level, the process could be driven by the import of compounds such as cytokinins transported in the transpiration stream (Pons et al., 2001; Boonman et al., 2007). Although the actual N distribution usually follows the light gradient, in all studies it is less steep than the calculated optimal N profile maximizing canopy photosynthesis (Pons et al., 1989; Yin et al., 2003). Possible reasons for this discrepancy have been discussed in detail by Kull (2002). Sink-source relations and in particular the demand for N could modulate the light-leaf N relationship (Dreccer et al., 1998), but conflicting results have been reported regarding the effect of N availability on the light-leaf N relationship. While some authors found no effect of N availability (Sinclair and Shiraiwa, 1993; Milroy et al., 2001), others found that the N gradient relative to light (i.e. KL/KN) was steeper under low N (Hikosaka et al., 1994; Grindlay et al., 1995; Lötscher et al., 2003) or that the response of the light-leaf N relationship to N availability depended on the developmental stage (Dreccer et al., 2000). Interspecific differences in the light-leaf N relationship have also been reported and were related to differences in phenotypic plasticity (Aerts, 1996) or plant architecture (leaf stature and branching pattern; Anten et al., 1995a; Lötscher et al., 2003).Since canopy photosynthesis is dependent upon the leaf N gradient, it has been suggested that the pattern of leaf N distribution could be responsible for part of the genetic variability associated with the negative correlation between grain yield and protein concentration reported for various crop species (Dreccer et al., 1998). In wheat (Triticum aestivum), N accumulated before anthesis contributes 30% to 70% of grain N (Mi et al., 2000; Kichey et al., 2007). The efficiency of N translocation from the lower to the upper leaves may increase with the steepness of the N gradient, with only a negligible effect on canopy carbon gain integrated over the whole grain-filling period. This hypothesis is consistent with experimental studies based on a range of genotypes showing that, at a given grain yield level, grain protein concentration is positively related to the efficiency of N translocation either from the lower to the upper leaves (Wang et al., 2005) or from the leaves to the grains (Monaghan et al., 2001; Jukanti and Fischer, 2008). Only a few studies have investigated the intraspecific variability of the light-N relationship at the intraspecific level (Shiraiwa and Sinclair, 1993; Bindraban, 1999; Bertheloot et al., 2008; van Oosterom et al., 2010). For wheat, published analyses of the genetic variability of the light-leaf N relationship were limited to only two to five genotypes, and no genetic differences were found (Bindraban, 1999; Bertheloot et al., 2008).This paper focuses on the genetic variability of the vertical leaf N gradient with respect to light for wheat. Three main issues were investigated. What is the effect of N supply on the vertical distribution of leaf N? Does the distribution of leaf N with respect to light differ among genotypes? If the adjustment of leaf N to the light gradient varies with both the genotype and N supply, could this genetic and environmental variability have a unique ecophysiological determinant (driving variable)?These questions were addressed using 16 genotypes (Supplemental Table S1) covering a wide range of variation for N use efficiency (i.e. grain dry mass yield per unit of available mineral N from the soil and fertilizer), for grain protein concentration (Le Gouis et al., 2000; Foulkes et al., 2006; Gaju et al., 2011) and for the deviation from the negative correlation between grain yield and protein concentration (Oury et al., 2003). The 16 genotypes were grown in the field under two conditions of N supply (N− and N+ for low- and high-N treatments, respectively) in order to modulate crop N status at Clermont-Ferrand (CF) in France in two consecutive seasons (experiments CF07 and CF08). In addition, four of the 16 cultivars representing the variability observed for N utilization and N uptake efficiency were grown in the field under two conditions of N supply at Sutton Bonington (SB) in the United Kingdom in one season (experiment SB07). The distribution of leaf N was analyzed at anthesis. The first reason for this is that the distribution of both light and leaf N within the canopy is relatively stable from this phenological stage until almost the end of grain filling (Bertheloot et al., 2008). Whereas the canopy green area index (GAI) decreases dramatically during the grain-filling period, the structure of the canopy affecting light interception does not change significantly during that period. Both the vertical light and N distributions down the canopy are unchanged during most of the grain-filling period; therefore, the KN-to-KL ratio is constant during that period (Bertheloot et al., 2008). Similarly, Archontoulis et al. (2011) showed that KN-to-KL ratio is not modified during the vegetative and reproductive stages for field-gown sunflower (Helianthus annuus) crops. Therefore, as most of the final grain yield results from carbon assimilated after anthesis (Bidinger et al., 1977; Gebbing and Schnyder, 1999), the N distribution at anthesis is very relevant in terms of carbon assimilation and grain yield in wheat. A second reason is that the number and potential size of grains are determined around anthesis, which therefore appears as a critical stage in the formation of grain yield. A better understanding of the ecophysiological determinants of leaf N gradient at this phenological stage could consequently be crucial for improving wheat productivity and quality (Dreccer et al., 1998).  相似文献   

2.
3.
4.
A major contributor to the global carbon cycle is plant respiration. Elevated atmospheric CO2 concentrations may either accelerate or decelerate plant respiration for reasons that have been uncertain. We recently established that elevated CO2 during the daytime decreases plant mitochondrial respiration in the light and protein concentration because CO2 slows the daytime conversion of nitrate (NO3) into protein. This derives in part from the inhibitory effect of CO2 on photorespiration and the dependence of shoot NO3 assimilation on photorespiration. Elevated CO2 also inhibits the translocation of nitrite into the chloroplast, a response that influences shoot NO3 assimilation during both day and night. Here, we exposed Arabidopsis (Arabidopsis thaliana) and wheat (Triticum aestivum) plants to daytime or nighttime elevated CO2 and supplied them with NO3 or ammonium as a sole nitrogen (N) source. Six independent measures (plant biomass, shoot NO3, shoot organic N, 15N isotope fractionation, 15NO3 assimilation, and the ratio of shoot CO2 evolution to O2 consumption) indicated that elevated CO2 at night slowed NO3 assimilation and thus decreased dark respiration in the plants reliant on NO3. These results provide a straightforward explanation for the diverse responses of plants to elevated CO2 at night and suggest that soil N source will have an increasing influence on the capacity of plants to mitigate human greenhouse gas emissions.The CO2 concentration in Earth’s atmosphere has increased from about 270 to 400 µmol mol–1 since 1800, and may double before the end of the century (Intergovernmental Panel on Climate Change, 2013). Plant responses to such increases are highly variable, but plant nitrogen (N) concentrations generally decline under elevated CO2 (Cotrufo et al., 1998; Long et al., 2004). One explanation for this decline is that CO2 inhibits nitrate (NO3) assimilation into protein in the shoots of C3 plants during the daytime (Bloom et al., 2002, 2010, 2012, 2014; Cheng et al., 2012; Pleijel and Uddling, 2012; Myers et al., 2014; Easlon et al., 2015; Pleijel and Högy, 2015). This derives in part from the inhibitory effect of CO2 on photorespiration (Foyer et al., 2009) and the dependence of shoot NO3 assimilation on photorespiration (Rachmilevitch et al., 2004; Bloom, 2015).A key factor in global carbon budgets is plant respiration at night (Amthor, 1991; Farrar and Williams, 1991; Drake et al., 1999; Leakey et al., 2009). Nighttime elevated CO2 may inhibit, have a negligible effect on, or stimulate dark respiration, depending on the plant species (Bunce, 2001, 2003; Wang and Curtis, 2002), plant development stage (Wang et al., 2001; Li et al., 2013), experimental approach (Griffin et al., 1999; Baker et al., 2000; Hamilton et al., 2001; Bruhn et al., 2002; Jahnke and Krewitt, 2002; Bunce, 2004), and total N supply (Markelz et al., 2014). The current study is, to our knowledge, the first to examine the influence of N source, NO3 versus ammonium (NH4+), on plant dark respiration at elevated CO2 during the night.Plant organic N compounds account for less than 5% of the total dry weight of a plant, but conversion of NO3 into organic N expends about 25% of the total energy in shoots (Bloom et al., 1989) and roots (Bloom et al., 1992). During the day, photorespiration supplies a portion of the energy (Rachmilevitch et al., 2004; Foyer et al., 2009), but at night, this energetic cost is borne entirely by the respiration of C substrates (Amthor, 1995) and may divert a substantial amount of reductant from the mitochondrial electron transport chain (Cousins and Bloom, 2004). The relative importance of NO3 assimilation at night versus the day, however, is still a matter of intense debate (Nunes-Nesi et al., 2010). Here, we estimated NO3 assimilation using several independent methods and show in Arabidopsis (Arabidopsis thaliana) and wheat (Triticum aestivum), two diverse C3 plants, that NO3 assimilation at night can be substantial, and that elevated CO2 at night inhibits this process.  相似文献   

5.
6.
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   

7.
8.
9.
When photosynthetic organisms are deprived of nitrogen (N), the capacity to grow and assimilate carbon becomes limited, causing a decrease in the productive use of absorbed light energy and likely a rise in the cellular reduction state. Although there is a scarcity of N in many terrestrial and aquatic environments, a mechanistic understanding of how photosynthesis adjusts to low-N conditions and the enzymes/activities integral to these adjustments have not been described. In this work, we use biochemical and biophysical analyses of photoautotrophically grown wild-type and mutant strains of Chlamydomonas reinhardtii to determine the integration of electron transport pathways critical for maintaining active photosynthetic complexes even after exposure of cells to N deprivation for 3 d. Key to acclimation is the type II NADPH dehydrogenase, NDA2, which drives cyclic electron flow (CEF), chlororespiration, and the generation of an H+ gradient across the thylakoid membranes. N deprivation elicited a doubling of the rate of NDA2-dependent CEF, with little contribution from PGR5/PGRL1-dependent CEF. The H+ gradient generated by CEF is essential to sustain nonphotochemical quenching, while an increase in the level of reduced plastoquinone would promote a state transition; both are necessary to down-regulate photosystem II activity. Moreover, stimulation of NDA2-dependent chlororespiration affords additional relief from the elevated reduction state associated with N deprivation through plastid terminal oxidase-dependent water synthesis. Overall, rerouting electrons through the NDA2 catalytic hub in response to photoautotrophic N deprivation sustains cell viability while promoting the dissipation of excess excitation energy through quenching and chlororespiratory processes.Oxygenic photosynthesis involves the conversion of light energy into chemical bond energy by plants, green algae, and cyanobacteria and the use of that energy to fix CO2. The photosynthetic electron transport system, located in thylakoid membranes, involves several major protein complexes: PSII (water-plastoquinone oxidoreductase), cytochrome b6f (cyt b6f; plastoquinone-plastocyanin oxidoreductase), PSI (plastocyanin-ferredoxin oxidoreductase), and the ATP synthase (CFoCF1). Light energy absorbed by the photosynthetic apparatus is used to establish both linear electron flow (LEF) and cyclic electron flow (CEF), which drive the production of ATP and NADPH, the chemical products of the light reactions needed for CO2 fixation in the Calvin-Benson-Bassham (CBB) cycle.With the absorption of light energy by pigment-protein complexes associated with PSII, energy is funneled into unique chlorophyll (Chl) molecules located in the PSII reaction center (RC), where it can elicit a charge separation that generates a large enough oxidizing potential to extract electrons from water. In LEF, electrons from PSII RCs are transferred sequentially along a set of electron carriers, initially reducing the plastoquinone (PQ) pool, then the cyt b6f complex, and subsequently the lumenal electron carrier plastocyanin (PC). Light energy absorbed by PSI excites a special pair of Chl molecules (P700), causing a charge separation that generates the most negative redox potential in nature (Nelson and Yocum, 2006). The energized electron, which is replaced by electrons from PC, is sequentially transferred to ferredoxin and ferredoxin NADP+ reductase, generating reductant in the form of NADPH.Electron transport from water to NADPH in LEF is accompanied by the transport of H+ into the thylakoid lumen. For each water molecule oxidized, two H+ are released in the thylakoid lumen. In addition, H+ are moved into the lumen by the transfer of electrons through cyt b6f (Q cycle). H+ accumulation in the thylakoid lumen dramatically alters the lumenal pH, and the transmembrane H+ gradient (ΔpH) together with the transmembrane ion gradient constitute the proton motive force (pmf), which drives ATP formation by ATP synthase (Mitchell, 1961, 1966, 2011). This pmf also promotes other cellular processes, including the dissipation of excess absorbed excitation energy as heat in a photoprotective process (see below; Li et al., 2009; Erickson et al., 2015). The NADPH and ATP molecules generated by LEF and CEF fuel the synthesis of reduced carbon backbones (in the CBB cycle) used in the production of many cellular metabolites and fixed carbon storage polymers.A basic role for CEF is to increase the ATP-NADPH ratio, which can satisfy the energy requirements of the cell and augment the synthesis of ATP by LEF, which is required to sustain CO2 fixation by the CBB cycle (Allen, 2003; Kramer et al., 2004; Iwai et al., 2010; Alric, 2014). There are two distinct CEF pathways identified in plants and algae. In both pathways, electrons flow from the PQ pool through cyt b6f to reduce the oxidized form of P700 (P700+). In one CEF pathway, electrons are transferred back to the PQ pool prior to the formation of NADPH. This route involves the proteins PGR5 and PGRL1 (DalCorso et al., 2008; Tolleter et al., 2011; Hertle et al., 2013) and is termed PGR5/L1-dependent CEF. A second route for CEF includes an NADPH dehydrogenase that oxidizes NADPH (product of LEF) to NADP+, simultaneously reducing the PQ (Allen, 2003; Kramer et al., 2004; Rumeau et al., 2007). The reduced PQ pool is then oxidized by cyt b6f, causing H+ translocation into the thylakoid lumen, followed by the transfer of electrons to P700+ via PC. In the green alga Chlamydomonas reinhardtii, this second route for CEF involves a type II NADPH dehydrogenase (NDA2; Jans et al., 2008; Desplats et al., 2009).Oxygenic photosynthetic organisms have inhabited the planet for approximately 3 billion years and have developed numerous strategies to acclimate to environmental fluctuations. These acclimation processes confer flexibility to the photosynthetic machinery, allowing it to adjust to changes in conditions that impact the metabolic/energetic state of the organism and, most importantly, the formation of reactive oxygen species that may damage the photosynthetic apparatus and other cellular components (Li et al., 2009). Several ways in which the photosynthetic apparatus adjusts to environmental fluctuations have been established. A well-studied acclimation process, nonphotochemical quenching (NPQ), reduces the excitation pressure on PSII when oxidized downstream electron acceptors are not available (Eberhard et al., 2008; Li et al., 2009; Erickson et al., 2015). Several processes constitute NPQ, as follows. (1) qT, which involves the physical movement of light-harvesting complexes (LHCs) from one photosystem to another (this is also designated state transitions; Rochaix, 2014). (2) qE, which involves thermal dissipation of the excitation energy. This energy-dependent process requires an elevated ΔpH and involves an LHC-like protein, LHCSR3 (in C. reinhardtii) or PSBS (in plants), as well as the accumulation of specific xanthophylls (mainly lutein in C. reinhardtii and zeaxanthin in plants; Niyogi et al., 1997b; Li et al., 2000, 2004; Peers et al., 2009). (3) qZ, which is energy independent and involves the accumulation of zeaxanthin (Dall’Osto et al., 2005; Nilkens et al., 2010). (4) qI, which promotes quenching following physical damage to PSII core subunits (Aro et al., 1993). Additional mechanisms that can impact LEF and CEF are the synthesis and degradation of pigment molecules, changes in levels of RC and antenna complexes, and the control of electron distribution between LEF and CEF as the energetic demands of the cell change (Allen, 2003; Kramer et al., 2004). In addition, electrons can be consumed by mitochondrial and chlororespiratory activities (Bennoun, 1982; Peltier and Cournac, 2002; Johnson et al., 2014; Bailleul et al., 2015). The latter mainly involves the plastid terminal oxidase PTOX2, which catalyzes the oxidation of the PQ pool and the reduction of oxygen and H+ to form water molecules (Houille-Vernes et al., 2011; Nawrocki et al., 2015).Photosynthetic processes also must be modulated as organisms experience changes in the levels of available nutrients (Grossman and Takahashi, 2001). The macronutrient nitrogen (N), which represents 3% to 5% of the dry weight of photosynthetic organisms, is required to synthesize many biological molecules (e.g. amino acids, nucleic acids, and various metabolites) and also participates in posttranslational modifications of proteins (e.g. S-nitrosylation; Romero-Puertas et al., 2013). Importantly, N is highly abundant in chloroplasts in the form of DNA, ribosomes, Chl, and polypeptides (e.g. Rubisco and LHCs; Evans, 1989; Raven, 2013). Furthermore, there is a strong integration between N and carbon assimilation. During N limitation under photoautotrophic conditions, the inability of the organism to synthesize amino acids and other N-containing molecules necessary for cell growth and division can feed back to inhibit both carbon fixation by the CBB cycle and electron transport processes and also can negatively impact the expression of genes encoding key CBB cycle enzymes (Terashima and Evans, 1988; Huppe and Turpin, 1994; Nunes-Nesi et al., 2010).C. reinhardtii is a well-established model organism in which to study photosynthesis and acclimation processes, including acclimation to nutrient limitation (Wykoff et al., 1998; Grossman and Takahashi, 2001; Moseley et al., 2006; Grossman et al., 2009; Terauchi et al., 2010; Aksoy et al., 2013). This unicellular alga grows rapidly as a photoheterotroph (on fixed carbon in the light) or as a heterotroph (on fixed carbon in the dark), has completely sequenced nuclear, chloroplast, and mitochondrial genomes, can be used for classical genetic analyses, and is haploid, which makes some aspects of molecular manipulation (e.g. the generation of knockout mutants) easier (Merchant et al., 2007; Blaby et al., 2014). In the past few years, there have been many studies on the ways in which C. reinhardtii responds to N deprivation (Bulté and Wollman, 1992; Blaby et al., 2013; Goodenough et al., 2014; Schmollinger et al., 2014; Wei et al., 2015; Juergens et al., 2015). Cells deprived of N under photoheterotrophic conditions (i.e. acetate as an external carbon source) minimize the use of N (referred to as N sparing) and induce mechanisms associated with scavenging N from both external and internal pools, all of which eventually lead to proteome modifications and an elevated carbon-N ratio (Schmollinger et al., 2014). Acclimation under photoheterotrophic conditions also causes dramatic modifications of cellular metabolism and energetics: photosynthesis is down-regulated at multiple levels, with a portion of its N content recycled (mainly Chl and polypeptides of the photosynthetic apparatus), while there is enhanced accumulation of mitochondrial complexes leading to increased respiratory activity (Schmollinger et al., 2014; Juergens et al., 2015). Additionally, while fixed carbon cannot be used for growth in the absence of N, it may be stored as starch and triacylglycerol (Work et al., 2010; Siaut et al., 2011; Davey et al., 2014; Goodenough et al., 2014).In contrast to the acclimation of photoheterotrophically grown C. reinhardtii to N deprivation, little is known about how the photosynthetic machinery in this alga adjusts in response to N deprivation under photoautotrophic conditions, when the cells absolutely require photosynthetic energy generation for maintenance. Specifically, we sought to understand how photosynthesis adjusts to metabolic restrictions that slow down the CBB cycle, which in turn could cause the accumulation of photoreductant, particularly NADPH, as the demand for electrons declines (Peltier and Schmidt, 1991; Rumeau et al., 2007). Based on analyses of mutants and the use of spectroscopic and fluorescence measurements, we established a critical role for NDA2 in the acclimation of C. reinhardtii to N deprivation under photoautotrophic conditions, including (1) an augmented capacity for alternative routes of electron utilization (which decrease the NADPH-NADP+ ratio) based on increased NDA2-dependent CEF and chlororespiration, and (2) elevated qE, which relies on the H+ gradient generated by NDA2-dependent CEF.  相似文献   

10.
11.
Suboptimal nitrogen (N) availability is a primary constraint for crop production in developing nations, while in rich nations, intensive N fertilization carries substantial environmental and economic costs. Therefore, understanding root phenes that enhance N acquisition is of considerable importance. Structural-functional modeling predicts that root cortical aerenchyma (RCA) could improve N acquisition in maize (Zea mays). We evaluated the utility of RCA for N acquisition by physiological comparison of maize recombinant inbred lines contrasting in RCA grown under suboptimal and adequate N availability in greenhouse mesocosms and in the field in the United States and South Africa. N stress increased RCA formation by 200% in mesocosms and by 90% to 100% in the field. RCA formation substantially reduced root respiration and root N content. Under low-N conditions, RCA formation increased rooting depth by 15% to 31%, increased leaf N content by 28% to 81%, increased leaf chlorophyll content by 22%, increased leaf CO2 assimilation by 22%, increased vegetative biomass by 31% to 66%, and increased grain yield by 58%. Our results are consistent with the hypothesis that RCA improves plant growth under N-limiting conditions by decreasing root metabolic costs, thereby enhancing soil exploration and N acquisition in deep soil strata. Although potential fitness tradeoffs of RCA formation are poorly understood, increased RCA formation appears be a promising breeding target for enhancing crop N acquisition.Nitrogen (N) deficiency is one of the most limiting factors in maize (Zea mays) production worldwide (Ladha et al., 2005). In developing countries such as those in sub-Saharan Africa, less than 20 kg N ha−1 is applied to fields of smallholder farmers due to high fertilizer cost (Azeez et al., 2006; Worku et al., 2007). In developed countries, intensive N fertilization is used to maintain satisfactory yield (Tilman et al., 2002). In the United States, N fertilizers are the greatest economic and energy cost for maize production (Ribaudo et al., 2011). However, less than half of the N applied to crops is actually acquired, and most of the remaining N becomes a source of environmental pollution (Raun and Johnson, 1999; Smil, 1999; Tilman et al., 2002). For example, N and phosphorus (P) effluents into marine systems from agriculture cause eutrophication and hypoxic zones (Diaz and Rosenberg, 2008; Robertson and Vitousek, 2009). Nitrate contamination in surface water and groundwater systems poses serious health risks, such as methemoglobinemia and N-nitroso-induced cancers (UNEP and WHRC, 2007). Emission of nitrous oxides from agricultural activities contributes to ozone damage and global warming (Kulkarni et al., 2008; Sutton et al., 2011). Furthermore, the production of N fertilizers requires considerable energy from fossil fuels, and since energy costs have risen in recent years, farmers face economic pressure from increasing N fertilizer costs, which are linked to higher food prices. It is estimated that a 1% increase in crop N efficiency could save more than $1 billion (U.S.) annually worldwide (Kant et al., 2011). Therefore, even a small improvement in N efficiency would have significant positive impacts on the environment and the economy.Soil N is heterogenous and dynamic. The bioavailability of soil N depends on the balance between the rates of mineralization, nitrification, and denitrification. These processes are determined by several factors, including soil composition, microbial activity, soil temperature, and soil water status (Miller and Cramer, 2004). The predominant form of soil N available to plants in most agricultural systems is nitrate, which is highly soluble in water and thus mobile in the soil (Barber, 1995; Marschner, 1995). Mineralization of organic matter and/or the application of N fertilizer at the beginning of the growing season followed by precipitation and irrigation create a pulse of nitrate that may exceed the N acquisition capacity of seedlings and leach below the root zone. Therefore, it has been proposed that increasing the speed of root exploration of deep soil strata could benefit N acquisition (Lynch, 2013). However, the structural investments and metabolic expenditures of root systems are substantial and can exceed half of daily photosynthesis (Lambers et al., 2002). Therefore, full consideration of the costs and benefits of root systems is crucial for identifying root traits to improve crop production, especially in water- and nutrient-deficient environments (Lynch, 2007). Taking rhizoeconomics and the spatiotemporal availability of soil N into account, Lynch (2013) proposed a root ideotype for enhanced N acquisition in maize called Steep, Cheap, and Deep, in which Steep refers to architectural phenes and Cheap refers to phenes that reduce the metabolic cost of soil exploration. One element of this ideotype is abundant root cortical aerenchyma (RCA).RCA consists of enlarged air spaces in the root cortex (Esau, 1977). RCA is known to form in response to hypoxia, and the role of RCA in improving oxygen transport to roots of many plant species under hypoxic conditions has been well researched (Vartapetian and Jackson, 1997; Jackson and Armstrong, 1999; Mano and Omori, 2007, 2013). Interestingly, RCA can also form in response to drought and edaphic stresses such as N, P, and sulfur deficiencies (Drew et al., 1989; Bouranis et al., 2003; Fan et al., 2003; Zhu et al., 2010a), which suggests that the benefit of RCA extends beyond facilitating oxygen transport. Several lines of evidence suggest that RCA enhances root metabolic efficiency under stress. Fan et al. (2003) found that RCA formation significantly reduced root segment respiration and P content of root tissue, which allowed greater shoot growth in soils with low P availability. Under drought, maize genotypes with high RCA formation had greater root length, deeper rooting, better leaf water status, and 8 times greater yield than closely related genotypes with low RCA (Zhu et al., 2010a). Effects of RCA on root respiration were more pronounced for large-diameter roots compared with small-diameter roots (Jaramillo et al., 2013). Results from the functional-structural plant model SimRoot showed that RCA formation could be an adaptive response to deficiency of N, P, and potassium by decreasing the metabolic cost of soil exploration. By reducing root respiration, RCA decreases the carbon cost of soil exploration, and by decreasing the N and P content of root tissue, RCA permits internal reallocation of nutrients to growing root tissue, which is particularly beneficial under conditions of low N and P availability (Postma and Lynch, 2011a). Under suboptimal P availability, RCA increased the growth of a simulated 40-d-old maize plant by 70% (Postma and Lynch, 2011b). In the case of N, RCA increased the growth of simulated maize plants up to 55% in low-N conditions, and plants benefit from RCA more in high-N-leaching environments than in low-N-leaching environments (Postma and Lynch, 2011a). In addition, the formation of RCA decreases critical soil nutrient levels, defined as the soil fertility below which growth is reduced, suggesting that cultivars with high RCA may require less fertilizer under nonstressed conditions. These in silico results suggest that RCA has potential utility for improving crop nutrient acquisition in both high- and low-input agroecosystems.The overall objective of this research was to assess the utility of RCA for N acquisition in maize under N-limiting conditions. Maize near-isophenic recombinant inbred lines (RILs) sharing a common genetic background (i.e. descending from the same parents) with common root phenotypes but contrasting in RCA formation were grown under N stress to test the hypothesis that RCA formation is associated with reduced root respiration, reduced tissue nutrient content, greater rooting depth, enhanced N acquisition, and therefore greater plant growth and yield under N limitation.  相似文献   

12.
The effects of water deficit on carbon and nitrogen metabolism were investigated in flag leaves of wild-type and transgenic rice (Oryza sativa japonica ‘Kitaake’) plants expressing ISOPENTENYLTRANSFERASE (IPT; encoding the enzyme that mediates the rate-limiting step in cytokinin synthesis) under the control of PSARK, a maturation- and stress-induced promoter. While the wild-type plants displayed inhibition of photosynthesis and nitrogen assimilation during water stress, neither carbon nor nitrogen assimilation was affected by stress in the transgenic PSARK::IPT plants. In the transgenic plants, photosynthesis was maintained at control levels during stress and the flag leaf showed increased sucrose (Suc) phosphate synthase activity and reduced Suc synthase and invertase activities, leading to increased Suc contents. The sustained carbon assimilation in the transgenic PSARK::IPT plants was well correlated with enhanced nitrate content, higher nitrate reductase activity, and sustained ammonium contents, indicating that the stress-induced cytokinin synthesis in the transgenic plants played a role in maintaining nitrate acquisition. Protein contents decreased and free amino acids increased in wild-type plants during stress, while protein content was preserved in the transgenic plants. Our results indicate that the stress-induced cytokinin synthesis in the transgenic plants promoted sink strengthening through a cytokinin-dependent coordinated regulation of carbon and nitrogen metabolism that facilitates an enhanced tolerance of the transgenic plants to water deficit.Plant hormones control many aspects of plant growth and development and the responses of plants to abiotic and biotic stresses. Cytokinins (CKs) have been shown to regulate plant cell differentiation, leaf senescence, and other key developmental processes (Sakakibara, 2006). It has also been shown that CKs regulate assimilate partitioning (Ronzhina and Mokronosov, 1994), sink strength (Kuiper, 1993), and source/sink relationships (Roitsch, 1999). The localized expression in tobacco (Nicotiana tabacum) of a promoterless ISOPENTENYLTRANSFERASE (IPT), a gene encoding the enzyme that catalyzes the rate-limiting step in CK synthesis, enhanced the local sink strength and quickly mobilized nutrients to the tissues with elevated CK (Guivarc’h et al., 2002). Changes in sink/source relationships were also observed in CK-deficient tobacco shoots and roots (Werner et al., 2008). Elevated CK levels enhanced the survival of plants under water-stress conditions (Rivero et al., 2007). The overexpression of IPT under the control of SENESCENCE-ASSOCIATED RECEPTOR KINASE (SARK; a maturation- and stress-induced promoter) improved the drought tolerance of both eudicots (Rivero et al., 2007; Qin et al., 2011) and monocots (Peleg et al., 2011). After a water-stress episode during the reproductive stages (pre and post anthesis), transgenic PSARK::IPT rice (Oryza sativa) plants displayed higher grain yield than the wild type (Peleg et al., 2011). The transgenic PSARK::IPT rice exhibited a differential expression of genes encoding enzymes associated with hormone synthesis and hormone-regulated pathways. These results suggested that changes in hormone homeostasis induced the modification of source/sink relationships in the transgenic plants, resulting in higher grain yields under stress conditions (Peleg et al., 2011).The maintenance of carbon (C) and nitrogen (N) assimilation is of paramount importance to ensure sink strength and improve stress tolerance without yield penalties. The interactions between C and N metabolism are vital for plant growth and development, and complex mechanisms operate in the plant to coordinate C assimilation with N metabolism (Nunes-Nesi et al., 2010). Thus, plants respond to changes in C and N metabolites through the regulation of translation and posttranslational modification mechanisms. C and N metabolites activate signaling pathways that regulate enzyme and transporter activities that control C and N fluxes, optimizing the plant response to developmental and environmental cues changing source/sink relationships (Coruzzi and Zhou, 2001).Plant hormones affect, either directly or indirectly, these pathways and can act antagonistically or synergistically when responding to environmental stress (Wilkinson et al., 2012). The exposure of plants to water-limiting conditions results in abscisic acid (ABA) synthesis that induces ABA-dependent gene expression (Yamaguchi-Shinozaki and Shinozaki, 2006), triggering the closure of stomata and reducing water loss during drought (Wilkinson and Davies, 2010). Other hormones, in particular CK, salicylic acid, ethylene, and jasmonic acid, also play direct or indirect roles in the plant responses to abiotic stress (Peleg and Blumwald, 2011). Under drought stress, plant CK content decreases, and the reduction in CK increases the plant responses to increasing ABA (Davies and Zhang, 1991), inducing stomata closure and inhibiting photosynthesis (Rivero et al., 2010). Our previous results suggested that the stress-induced CK synthesis, driven by a stress-induced promoter, protected against the deleterious effects of water deficit on the photosynthetic apparatus, allowing higher photosynthetic rates and higher yields after water deficit in tobacco (Rivero et al., 2009) and cotton (Gossypium hirsutum; Kuppu et al., 2013) plants grown in the greenhouse and peanut (Arachis hypogaea) plants grown under field conditions (Qin et al., 2011).Here, we analyzed gene expression profiles, metabolites, and enzymatic and photosynthetic activities of flag leaves of wild-type and transgenic rice expressing PSARK::IPT exposed to water deficit during the reproductive stage and identified metabolic processes associated with the enhanced tolerance of the transgenic plants to water deficit. Our results indicate that the stress-induced CK synthesis in the transgenic plants promoted sink strengthening through the maintenance and coordination of N and C assimilation during water stress.  相似文献   

13.
Leaf shrinkage with dehydration has attracted attention for over 100 years, especially as it becomes visibly extreme during drought. However, little has been known of its correlation with physiology. Computer simulations of the leaf hydraulic system showed that a reduction of hydraulic conductance of the mesophyll pathways outside the xylem would cause a strong decline of leaf hydraulic conductance (Kleaf). For 14 diverse species, we tested the hypothesis that shrinkage during dehydration (i.e. in whole leaf, cell and airspace thickness, and leaf area) is associated with reduction in Kleaf at declining leaf water potential (Ψleaf). We tested hypotheses for the linkage of leaf shrinkage with structural and physiological water relations parameters, including modulus of elasticity, osmotic pressure at full turgor, turgor loss point (TLP), and cuticular conductance. Species originating from moist habitats showed substantial shrinkage during dehydration before reaching TLP, in contrast with species originating from dry habitats. Across species, the decline of Kleaf with mild dehydration (i.e. the initial slope of the Kleaf versus Ψleaf curve) correlated with the decline of leaf thickness (the slope of the leaf thickness versus Ψleaf curve), as expected based on predictions from computer simulations. Leaf thickness shrinkage before TLP correlated across species with lower modulus of elasticity and with less negative osmotic pressure at full turgor, as did leaf area shrinkage between full turgor and oven desiccation. These findings point to a role for leaf shrinkage in hydraulic decline during mild dehydration, with potential impacts on drought adaptation for cells and leaves, influencing plant ecological distributions.As leaves open their stomata to capture CO2 for photosynthesis, water is lost to transpiration, which needs to be replaced by flow through the hydraulic system. The leaf hydraulic system has two components, which act essentially in series: the pathways for water movement through the xylem from the petiole to leaf minor veins, and those through the living bundle sheath and mesophyll cells to the sites of evaporation (Tyree and Zimmermann, 2002; Sack et al., 2004; Sack and Holbrook, 2006). The decline in leaf hydraulic conductance (Kleaf) with dehydration may thus depend on both components. The importance of the xylem component is well established. Vein xylem embolism and cell collapse have been observed in dehydrating leaves (Salleo et al., 2001; Cochard et al., 2004a; Johnson et al., 2009), and computer modeling and experimental work showed that species with high major vein length per leaf area (VLA; i.e. for the first three vein-branching orders) were more resistant to hydraulic decline, providing more pathways around embolisms (Scoffoni et al., 2011). However, the physical impacts of dehydration on the extraxylem pathways have not been studied, even though in turgid leaves these pathways account for 26% to 88% of leaf hydraulic resistance (i.e. of 1/Kleaf), depending on species (Sack et al., 2003a; Cochard et al., 2004b). The aim of this study was to determine whether leaf shrinkage during dehydration relates to the decline of Kleaf as well as the structural determinants of leaf shrinkage.The shrinkage of leaves with dehydration has drawn attention for over 100 years. Leaves shrink in their area (Bogue, 1892; Gardner and Ehlig, 1965; Jones, 1973; Tang and Boyer, 2007; Blonder et al., 2012) and, considered in relative terms, even more strongly in their thickness (Fig. 1; Meidner, 1952; Gardner and Ehlig, 1965; Downey and Miller, 1971; Syvertsen and Levy, 1982; Saini and Rathore, 1983; Burquez, 1987; McBurney, 1992; Sancho-Knapik et al., 2010, 2011). Leaves fluctuate in thickness daily and seasonally according to transpiration (Kadoya et al., 1975; Tyree and Cameron, 1977; Fensom and Donald, 1982; Rozema et al., 1987; Ogaya and Peñuelas, 2006; Seelig et al., 2012). Indeed, the relation of leaf thickness to water status is so tight that using leaf thickness to guide irrigation has led to water savings of up to 45% (Seelig et al., 2012).Open in a separate windowFigure 1.Sketches of a fully turgid leaf (A) versus a strongly dehydrated leaf (B; drawings based on leaf cross sections of sunflower in Fellows and Boyer, 1978). Note the strong reduction in leaf thickness, cell thickness, and intercellular airspaces in the dehydrated leaf. Epidermal cells are shrunk in the dehydrated leaf, inducing whole-leaf area shrinkage. Note that this sketch represents shrinkage for a typical drought-sensitive species. Many species such as oaks (Quercus spp.) will experience less thickness shrinkage and an increase in intercellular airspace (see “Discussion”). [See online article for color version of this figure.]Previous studies of leaf shrinkage with progressive dehydration have tended to focus on single or few species. These studies showed that thickness declines with water status in two phases. Before the bulk leaf turgor loss point (TLP; leaf water potential [Ψleaf] at TLP) is reached, the slope of leaf thickness versus Ψleaf or relative water content (RWC) is shallower than past TLP for most species (Meidner, 1955, Kennedy and Booth, 1958, Burquez, 1987, McBurney, 1992, Sancho-Knapik et al., 2010, 2011). This is because before TLP, declining Ψleaf is strongly driven by declines in turgor pressure, which have a relatively low impact on cell and airspace volume, whereas past the TLP, declining Ψleaf depends only on solute concentration, which increases in inverse proportion as cell water volume declines while airspaces may shrink or expand (Tyree and Hammel, 1972, Sancho-Knapik et al., 2011). However, the steepness of the slope of leaf thickness versus Ψleaf before TLP seems to vary strongly across species (Meidner, 1955; Kennedy and Booth, 1958; Fellows and Boyer, 1978; Burquez, 1987; Colpitts and Coleman, 1997; Sancho-Knapik et al., 2010).A high leaf cell volume and turgor is crucial to physiological processes (Boyer, 1968; Lawlor and Cornic, 2002). Shrinkage may affect cell connectivity and water transport (Sancho-Knapik et al., 2011). However, no studies have tested for a possible relationship of leaf shrinkage with the decline of Kleaf during dehydration. Such an association would arise if, across species, shrinkage occurred simultaneously with vein xylem embolism or if tissue shrinkage led to declines in the extraxylem hydraulic conductance.To refine our hypotheses, we modified a computer model of the leaf hydraulic system (Cochard et al., 2004b; McKown et al., 2010; Scoffoni et al., 2011) to predict the impact of losses of xylem and extraxylem conductance on the response of Kleaf to dehydration. We characterized the degree of leaf shrinkage in thickness, in the thickness of cells and airspaces within the leaf, and in leaf area for 14 species diverse in phylogeny, leaf traits, and drought tolerance. We hypothesized that loss of extraxylem hydraulic conductance should have a greater impact on Kleaf at less negative water potentials when xylem tensions are too weak to trigger embolism and induce dramatic declines in Kleaf. We hypothesized that species with greater degrees of shrinkage before TLP would experience greater loss of Kleaf. Furthermore, we hypothesized that species from moist habitats would have greater degrees of shrinkage.For insight into the mechanisms and consequences of leaf shrinkage, we also investigated the relationships of 18 indices of leaf shrinkage with a wide range of aspects of leaf structure and composition, including gross morphology, leaf venation architecture, parameters of pressure-volume curves, and leaf water storage. We hypothesized that, across species, shrinkage in whole leaf, cell, and intercellular airspace thickness would be lower for species with greater allocation to structural rigidity and osmotic concentration, and thus shrinkage would be positively correlated with a lower modulus of elasticity (ε), less negative osmotic pressure at full turgor (πo), lower leaf mass per area (LMA), and lower leaf density. Additionally, we tested the longstanding hypothesis that species with higher major VLA and/or minor VLA (i.e. the fourth and higher vein-branching orders) would shrink less in area and/or thickness with dehydration (Gardner and Ehlig, 1965). Finally, we tested the ability of dehydrated leaves to recover in size with rehydration. We hypothesized that recovery would be greater for mildly than for strongly dehydrated leaves and that species with greater leaf shrinkage would be better able to recover from shrinkage.  相似文献   

14.
15.
The objective of this study was to determine if low stomatal conductance (g) increases growth, nitrate (NO3) assimilation, and nitrogen (N) utilization at elevated CO2 concentration. Four Arabidopsis (Arabidopsis thaliana) near isogenic lines (NILs) differing in g were grown at ambient and elevated CO2 concentration under low and high NO3 supply as the sole source of N. Although g varied by 32% among NILs at elevated CO2, leaf intercellular CO2 concentration varied by only 4% and genotype had no effect on shoot NO3 concentration in any treatment. Low-g NILs showed the greatest CO2 growth increase under N limitation but had the lowest CO2 growth enhancement under N-sufficient conditions. NILs with the highest and lowest g had similar rates of shoot NO3 assimilation following N deprivation at elevated CO2 concentration. After 5 d of N deprivation, the lowest g NIL had 27% lower maximum carboxylation rate and 23% lower photosynthetic electron transport compared with the highest g NIL. These results suggest that increased growth of low-g NILs under N limitation most likely resulted from more conservative N investment in photosynthetic biochemistry rather than from low g.The availability of water varies in time and space, and plants in a given environment are expected to evolve a stomatal behavior that optimizes the tradeoff of CO2 uptake for photosynthesis at the cost of transpirational water loss. The resource of CO2 also varies over time, and plant fossils indicate that stomatal characteristics have changed in response to periods of high and low atmospheric CO2 over the past 65 million years (Beerling and Chaloner, 1993; Van Der Burgh et al., 1993; Beerling, 1998; Kürschner, 2001; Royer et al., 2001). Relatively low atmospheric CO2 concentrations (less than 320 µmol mol−1) over the last 23 million years (Pearson and Palmer, 2000) are associated with increased stomatal conductance (g) to avoid CO2 starvation (Beerling and Chaloner, 1993). Atmospheric CO2 concentration has risen rapidly from 280 to 400 µmol mol−1 since 1800 and has resulted in lower stomatal density (Woodward, 1987; Woodward and Bazzaz, 1988; Lammertsma et al., 2011). At the current atmospheric CO2 concentration (400 µmol mol−1), further decreases in g reduce water loss but also restrict CO2 assimilation and, thus, limit the effectiveness of low g in water-stressed environments (Comstock and Ehleringer, 1993; Virgona and Farquhar, 1996). Elevated CO2 concentration enhances the diffusion gradient for CO2 into leaves, which allows g to decrease without severely restricting photosynthetic carbon gain (Herrick et al., 2004). Most consider such an improvement in water use efficiency in C3 plants to be the main driving force for decreased g at elevated CO2 concentration, especially in dry environments (Woodward, 1987; Beerling and Chaloner, 1993; Brodribb et al., 2009; Franks and Beerling, 2009; Katul et al., 2010).Water is the most common factor limiting terrestrial plant productivity, but declining stomatal density has also occurred in wetland environments where water stress is uncommon (Wagner et al., 2005). Improved water use efficiency at elevated CO2 concentration may be shifting the most common factor limiting plant productivity from water to nitrogen (N). In herbarium specimens of 14 species of trees, shrubs, and herbs, leaf N decreased 31% as atmospheric CO2 increased from about 270 to 400 μmol mol−1 since 1750 (Penuelas and Matamala, 1990). Indeed, many studies have shown that N availability limits the stimulation of plant growth at elevated CO2 concentration (Luo et al., 2004; Dukes et al., 2005; Reich et al., 2006). That most plants at elevated CO2 concentration exhibit both lower g and greater N limitation suggests a relationship between these factors.Plants primarily absorb N as nitrate (NO3) in most temperate soils and assimilate a major portion of this NO3 in shoots (Epstein and Bloom, 2005). Elevated CO2 increases the ratio of CO2 to oxygen in the chloroplast, decreasing photorespiration and improving photosynthetic efficiency (Sharkey, 1988) but inhibiting photorespiration-dependent NO3 assimilation (Rachmilevitch et al., 2004; Bloom et al., 2010, 2012; Bloom, 2014). Greater rhizosphere NO3 availability tends to enhance root NO3 assimilation and decrease the influence of elevated CO2 concentration on plant organic N accumulation (Kruse et al., 2002, 2003; Bloom et al., 2010).The most important factor regulating chloroplast CO2 concentration among natural accessions of Arabidopsis (Arabidopsis thaliana) is g and to a lesser extent mesophyll conductance (Easlon et al., 2014). Low g may decrease the ratio of CO2 to oxygen in the chloroplast at elevated CO2 concentration, enhancing photorespiration-dependent NO3 assimilation. Alternatively, increasing atmospheric CO2 may down-regulate the need to synthesize enzymes such as Rubisco to support photosynthesis, which conserves organic N, and g may decline as a by-product of lower photosynthetic capacity (Sage et al., 1989; Moore et al., 1998).Here, we examined the influence of atmospheric CO2 concentration and NO3 supply on photosynthesis, leaf N, and growth in near isogenic lines (NILs) of Arabidopsis differing in g. Arabidopsis accessions differ in many traits (including g) and likewise differ in DNA sequence at a large percentage of genes across the genome (Cao et al., 2011). Use of these NILs greatly reduces the proportion of the genome that varies and minimizes the influence of variation in other traits that are frequently associated with low g and could limit growth (Arp et al., 1998). We tested the extent to which (1) low g was associated with greater CO2 growth enhancement at low and high NO3 supply; (2) low leaf intercellular CO2 concentration (Ci) increased shoot NO3 assimilation; and (3) low g at elevated CO2 concentration was associated with altered N utilization in photosynthetic biochemistry.  相似文献   

16.
Necrotrophic and biotrophic pathogens are resisted by different plant defenses. While necrotrophic pathogens are sensitive to jasmonic acid (JA)-dependent resistance, biotrophic pathogens are resisted by salicylic acid (SA)- and reactive oxygen species (ROS)-dependent resistance. Although many pathogens switch from biotrophy to necrotrophy during infection, little is known about the signals triggering this transition. This study is based on the observation that the early colonization pattern and symptom development by the ascomycete pathogen Plectosphaerella cucumerina (P. cucumerina) vary between inoculation methods. Using the Arabidopsis (Arabidopsis thaliana) defense response as a proxy for infection strategy, we examined whether P. cucumerina alternates between hemibiotrophic and necrotrophic lifestyles, depending on initial spore density and distribution on the leaf surface. Untargeted metabolome analysis revealed profound differences in metabolic defense signatures upon different inoculation methods. Quantification of JA and SA, marker gene expression, and cell death confirmed that infection from high spore densities activates JA-dependent defenses with excessive cell death, while infection from low spore densities induces SA-dependent defenses with lower levels of cell death. Phenotyping of Arabidopsis mutants in JA, SA, and ROS signaling confirmed that P. cucumerina is differentially resisted by JA- and SA/ROS-dependent defenses, depending on initial spore density and distribution on the leaf. Furthermore, in situ staining for early callose deposition at the infection sites revealed that necrotrophy by P. cucumerina is associated with elevated host defense. We conclude that P. cucumerina adapts to early-acting plant defenses by switching from a hemibiotrophic to a necrotrophic infection program, thereby gaining an advantage of immunity-related cell death in the host.Plant pathogens are often classified as necrotrophic or biotrophic, depending on their infection strategy (Glazebrook, 2005; Nishimura and Dangl, 2010). Necrotrophic pathogens kill living host cells and use the decayed plant tissue as a substrate to colonize the plant, whereas biotrophic pathogens parasitize living plant cells by employing effector molecules that suppress the host immune system (Pel and Pieterse, 2013). Despite this binary classification, the majority of pathogenic microbes employ a hemibiotrophic infection strategy, which is characterized by an initial biotrophic phase followed by a necrotrophic infection strategy at later stages of infection (Perfect and Green, 2001). The pathogenic fungi Magnaporthe grisea, Sclerotinia sclerotiorum, and Mycosphaerella graminicola, the oomycete Phytophthora infestans, and the bacterial pathogen Pseudomonas syringae are examples of hemibiotrophic plant pathogens (Perfect and Green, 2001; Koeck et al., 2011; van Kan et al., 2014; Kabbage et al., 2015).Despite considerable progress in our understanding of plant resistance to necrotrophic and biotrophic pathogens (Glazebrook, 2005; Mengiste, 2012; Lai and Mengiste, 2013), recent debate highlights the dynamic and complex interplay between plant-pathogenic microbes and their hosts, which is raising concerns about the use of infection strategies as a static tool to classify plant pathogens. For instance, the fungal genus Botrytis is often labeled as an archetypal necrotroph, even though there is evidence that it can behave as an endophytic fungus with a biotrophic lifestyle (van Kan et al., 2014). The rice blast fungus Magnaporthe oryzae, which is often classified as a hemibiotrophic leaf pathogen (Perfect and Green, 2001; Koeck et al., 2011), can adopt a purely biotrophic lifestyle when infecting root tissues (Marcel et al., 2010). It remains unclear which signals are responsible for the switch from biotrophy to necrotrophy and whether these signals rely solely on the physiological state of the pathogen, or whether host-derived signals play a role as well (Kabbage et al., 2015).The plant hormones salicylic acid (SA) and jasmonic acid (JA) play a central role in the activation of plant defenses (Glazebrook, 2005; Pieterse et al., 2009, 2012). The first evidence that biotrophic and necrotrophic pathogens are resisted by different immune responses came from Thomma et al. (1998), who demonstrated that Arabidopsis (Arabidopsis thaliana) genotypes impaired in SA signaling show enhanced susceptibility to the biotrophic pathogen Hyaloperonospora arabidopsidis (formerly known as Peronospora parastitica), while JA-insensitive genotypes were more susceptible to the necrotrophic fungus Alternaria brassicicola. In subsequent years, the differential effectiveness of SA- and JA-dependent defense mechanisms has been confirmed in different plant-pathogen interactions, while additional plant hormones, such as ethylene, abscisic acid (ABA), auxins, and cytokinins, have emerged as regulators of SA- and JA-dependent defenses (Bari and Jones, 2009; Cao et al., 2011; Pieterse et al., 2012). Moreover, SA- and JA-dependent defense pathways have been shown to act antagonistically on each other, which allows plants to prioritize an appropriate defense response to attack by biotrophic pathogens, necrotrophic pathogens, or herbivores (Koornneef and Pieterse, 2008; Pieterse et al., 2009; Verhage et al., 2010).In addition to plant hormones, reactive oxygen species (ROS) play an important regulatory role in plant defenses (Torres et al., 2006; Lehmann et al., 2015). Within minutes after the perception of pathogen-associated molecular patterns, NADPH oxidases and apoplastic peroxidases generate early ROS bursts (Torres et al., 2002; Daudi et al., 2012; O’Brien et al., 2012), which activate downstream defense signaling cascades (Apel and Hirt, 2004; Torres et al., 2006; Miller et al., 2009; Mittler et al., 2011; Lehmann et al., 2015). ROS play an important regulatory role in the deposition of callose (Luna et al., 2011; Pastor et al., 2013) and can also stimulate SA-dependent defenses (Chaouch et al., 2010; Yun and Chen, 2011; Wang et al., 2014; Mammarella et al., 2015). However, the spread of SA-induced apoptosis during hyperstimulation of the plant immune system is contained by the ROS-generating NADPH oxidase RBOHD (Torres et al., 2005), presumably to allow for the sufficient generation of SA-dependent defense signals from living cells that are adjacent to apoptotic cells. Nitric oxide (NO) plays an additional role in the regulation of SA/ROS-dependent defense (Trapet et al., 2015). This gaseous molecule can stimulate ROS production and cell death in the absence of SA while preventing excessive ROS production at high cellular SA levels via S-nitrosylation of RBOHD (Yun et al., 2011). Recently, it was shown that pathogen-induced accumulation of NO and ROS promotes the production of azelaic acid, a lipid derivative that primes distal plants for SA-dependent defenses (Wang et al., 2014). Hence, NO, ROS, and SA are intertwined in a complex regulatory network to mount local and systemic resistance against biotrophic pathogens. Interestingly, pathogens with a necrotrophic lifestyle can benefit from ROS/SA-dependent defenses and associated cell death (Govrin and Levine, 2000). For instance, Kabbage et al. (2013) demonstrated that S. sclerotiorum utilizes oxalic acid to repress oxidative defense signaling during initial biotrophic colonization, but it stimulates apoptosis at later stages to advance necrotrophic colonization. Moreover, SA-induced repression of JA-dependent resistance not only benefits necrotrophic pathogens but also hemibiotrophic pathogens after having switched from biotrophy to necrotrophy (Glazebrook, 2005; Pieterse et al., 2009, 2012).Plectosphaerella cucumerina ((P. cucumerina, anamorph Plectosporum tabacinum) anamorph Plectosporum tabacinum) is a filamentous ascomycete fungus that can survive saprophytically in soil by decomposing plant material (Palm et al., 1995). The fungus can cause sudden death and blight disease in a variety of crops (Chen et al., 1999; Harrington et al., 2000). Because P. cucumerina can infect Arabidopsis leaves, the P. cucumerina-Arabidopsis interaction has emerged as a popular model system in which to study plant defense reactions to necrotrophic fungi (Berrocal-Lobo et al., 2002; Ton and Mauch-Mani, 2004; Carlucci et al., 2012; Ramos et al., 2013). Various studies have shown that Arabidopsis deploys a wide range of inducible defense strategies against P. cucumerina, including JA-, SA-, ABA-, and auxin-dependent defenses, glucosinolates (Tierens et al., 2001; Sánchez-Vallet et al., 2010; Gamir et al., 2014; Pastor et al., 2014), callose deposition (García-Andrade et al., 2011; Gamir et al., 2012, 2014; Sánchez-Vallet et al., 2012), and ROS (Tierens et al., 2002; Sánchez-Vallet et al., 2010; Barna et al., 2012; Gamir et al., 2012, 2014; Pastor et al., 2014). Recent metabolomics studies have revealed large-scale metabolic changes in P. cucumerina-infected Arabidopsis, presumably to mobilize chemical defenses (Sánchez-Vallet et al., 2010; Gamir et al., 2014; Pastor et al., 2014). Furthermore, various chemical agents have been reported to induce resistance against P. cucumerina. These chemicals include β-amino-butyric acid, which primes callose deposition and SA-dependent defenses, benzothiadiazole (BTH or Bion; Görlach et al., 1996; Ton and Mauch-Mani, 2004), which activates SA-related defenses (Lawton et al., 1996; Ton and Mauch-Mani, 2004; Gamir et al., 2014; Luna et al., 2014), JA (Ton and Mauch-Mani, 2004), and ABA, which primes ROS and callose deposition (Ton and Mauch-Mani, 2004; Pastor et al., 2013). However, among all these studies, there is increasing controversy about the exact signaling pathways and defense responses contributing to plant resistance against P. cucumerina. While it is clear that JA and ethylene contribute to basal resistance against the fungus, the exact roles of SA, ABA, and ROS in P. cucumerina resistance vary between studies (Thomma et al., 1998; Ton and Mauch-Mani, 2004; Sánchez-Vallet et al., 2012; Gamir et al., 2014).This study is based on the observation that the disease phenotype during P. cucumerina infection differs according to the inoculation method used. We provide evidence that the fungus follows a hemibiotrophic infection strategy when infecting from relatively low spore densities on the leaf surface. By contrast, when challenged by localized host defense to relatively high spore densities, the fungus switches to a necrotrophic infection program. Our study has uncovered a novel strategy by which plant-pathogenic fungi can take advantage of the early immune response in the host plant.  相似文献   

17.
18.
Two mutants sensitive to heat stress for growth and impaired in NADPH dehydrogenase (NDH-1)-dependent cyclic electron transport around photosystem I (NDH-CET) were isolated from the cyanobacterium Synechocystis sp. strain PCC 6803 transformed with a transposon-bearing library. Both mutants had a tag in the same sll0272 gene, encoding a protein highly homologous to NdhV identified in Arabidopsis (Arabidopsis thaliana). Deletion of the sll0272 gene (ndhV) did not influence the assembly of NDH-1 complexes and the activities of CO2 uptake and respiration but reduced the activity of NDH-CET. NdhV interacted with NdhS, a ferredoxin-binding subunit of cyanobacterial NDH-1 complex. Deletion of NdhS completely abolished NdhV, but deletion of NdhV had no effect on the amount of NdhS. Reduction of NDH-CET activity was more significant in ΔndhS than in ΔndhV. We therefore propose that NdhV cooperates with NdhS to accept electrons from reduced ferredoxin.Cyanobacterial NADPH dehydrogenase (NDH-1) complexes are localized in the thylakoid membrane (Ohkawa et al., 2001, 2002; Zhang et al., 2004; Xu et al., 2008; Battchikova et al., 2011b) and participate in a variety of bioenergetic reactions, such as respiration, cyclic electron transport around photosystem I (NDH-CET), and CO2 uptake (Ogawa, 1991; Mi et al., 1992; Ohkawa et al., 2000). Structurally, the cyanobacterial NDH-1 complexes closely resemble energy-converting complex I in eubacteria and the mitochondrial respiratory chain regardless of the absence of homologs of three subunits in cyanobacterial genomes that constitute the catalytically active core of complex I (Friedrich et al., 1995; Friedrich and Scheide, 2000; Arteni et al., 2006). Over the past decade, new subunits of NDH-1 complexes specific to oxygenic photosynthesis have been identified in several cyanobacterial strains. They are NdhM to NdhQ and NdhS (Prommeenate et al., 2004; Battchikova et al., 2005, 2011b; Nowaczyk et al., 2011; Wulfhorst et al., 2014; Zhang et al., 2014; Zhao et al., 2014b, 2015), in addition to NdhL first identified in the cyanobacterium Synechocystis sp. strain PCC 6803 (hereafter Synechocystis 6803) about 20 years ago (Ogawa, 1992). Among them, NdhS possesses a ferredoxin (Fd)-binding motif and was shown to bind Fd, which suggested that Fd is one of the electron donors to NDH-1 complexes (Mi et al., 1995; Battchikova et al., 2011b; Ma and Ogawa, 2015). Deletion of NdhS strongly reduced the activity of NDH-CET but had no effect on respiration and CO2 uptake (Battchikova et al., 2011b; Ma and Ogawa, 2015). The NDH-CET plays an important role in coping with various environmental stresses regardless of its elusive mechanism. For example, this function can greatly alleviate heat-sensitive growth phenotypes (Wang et al., 2006a; Zhao et al., 2014a). Thus, heat treatment strategy can help in identifying the proteins essential to NDH-CET.Here, a new oxygenic photosynthesis-specific (OPS) subunit NdhV was identified in Synechocystis 6803 with the help of heat treatment strategy, and its deletion did not influence the assembly of NDH-1L and NDH-1MS complexes and the activities of CO2 uptake and respiration but impaired the NDH-CET activity. We give evidence that NdhV interacts with NdhS and is another component of Fd-binding domain of cyanobacterial NDH-1 complex. A possible role of NdhV on the NDH-CET activity is discussed.  相似文献   

19.
Stomatal responses to vapor pressure deficit (VPD) are a principal means by which vascular land plants regulate daytime transpiration. While much work has focused on characterizing and modeling this response, there remains no consensus as to the mechanism that drives it. Explanations range from passive regulation by leaf hydration to biochemical regulation by the phytohormone abscisic acid (ABA). We monitored ABA levels, leaf gas exchange, and water status in a diversity of vascular land plants exposed to a symmetrical, mild transition in VPD. The stomata in basal lineages of vascular plants, including gymnosperms, appeared to respond passively to changes in leaf water status induced by VPD perturbation, with minimal changes in foliar ABA levels and no hysteresis in stomatal action. In contrast, foliar ABA appeared to drive the stomatal response to VPD in our angiosperm samples. Increased foliar ABA level at high VPD in angiosperm species resulted in hysteresis in the recovery of stomatal conductance; this was most pronounced in herbaceous species. Increased levels of ABA in the leaf epidermis were found to originate from sites of synthesis in other parts of the leaf rather than from the guard cells themselves. The transition from a passive regulation to ABA regulation of the stomatal response to VPD in the earliest angiosperms is likely to have had critical implications for the ecological success of this lineage.Plants continuously regulate transpiration by controlling the aperture of the stomatal pores on the surface of the leaf. The principal atmospheric determinant of stomatal aperture is the humidity of the air, which can be expressed as the vapor pressure difference between the leaf and the atmosphere. Stomatal responses to atmospheric vapor pressure deficit (VPD) have been well characterized across the diversity of vascular plant species (Darwin, 1898; Lange et al., 1971; Turner et al., 1984; Franks and Farquhar, 1999; Oren et al., 1999; Brodribb and McAdam, 2011; Mott and Peak, 2013), with stomata typically closing at high VPD and opening at low VPD. This comprehensive characterization has allowed for the development of highly effective empirical and mechanistic models of leaf gas exchange that provide robust predictions of the responses of transpiration to changes in VPD (Buckley et al., 2003; Katul et al., 2009; Damour et al., 2010; Medlyn et al., 2011). Despite the success of this modeling, the mechanism for the stomatal response to VPD remains poorly understood (Damour et al., 2010). Different hypotheses range from one extreme, whereby stomata respond passively through changes in leaf water content induced by the VPD or humidity perturbation (Lange et al., 1971; Mott and Peak, 2013), to the other extreme, whereby stomata close uniquely in response to the phytohormone abscisic acid (ABA; Xie et al., 2006; Bauer et al., 2013).From the earliest recognition that stomata open and close by changes in guard cell turgor (Heath, 1938), there have been many attempts to link the passive changes in water status that occur during VPD or humidity transitions with stomatal responses to VPD or humidity (Lange et al., 1971; Mott and Peak, 2013). Studies have suggested that changes in atmospheric water content passively drive stomatal responses by changing bulk leaf water status, which in turn changes guard cell turgor (Oren et al., 1999), or alternatively by changing guard cell turgor directly (Mott and Peak, 2013). Models based on these entirely passive processes are highly effective in predicting steady-state stomatal conductance (gs) in response to changes in VPD or humidity in angiosperms (Mott and Peak, 2013).While hydraulic models provide robust predictions of steady-state gs, they are less effective at predicting the dynamic responses of stomata to short-term perturbations, particularly with respect to the wrong-way responses that typically occur as transients (Buckley, 2005), as well as feed-forward behavior (Farquhar, 1978; Bunce, 1997; Franks et al., 1997; Tardieu and Simonneau, 1998; Ocheltree et al., 2014; compare with Mott and Peak, 2013). Although some of these models provide a pathway for incorporating the effect of ABA (Buckley, 2005), a lack of knowledge of ABA dynamics or action makes it difficult to integrate the influence of this active regulator of guard cell aperture into models. The stomatal behavior of single gene mutants (most notably the ABA synthesis and signaling mutants of Arabidopsis) strongly supports a role for ABA in mediating standard stomatal responses to changes in VPD. The stomata of these mutants are known to have less pronounced responses to a reduction in relative humidity compared with wild-type plants (Xie et al., 2006). Recently, molecular work has shown that guard cells express many of the genes required to synthesize ABA (Okamoto et al., 2009; Bauer et al., 2013), with molecular proxies for ABA level also indicating that the biochemical activity of ABA in the guard cell may increase following short-term exposure of leaves to a reduction in relative humidity (Waadt et al., 2014). These findings suggest a role for ABA in regulating stomatal responses to VPD and have led some to the conclusion that ABA synthesized autonomously by the guard cells is the predominant mechanism for stomatal responses to increased VPD (Bauer et al., 2013).Although the experimental evidence from molecular studies presents an argument for the role of ABA in the responses of stomata to changes in VPD, very few studies have quantified changes in ABA level in response to VPD. It is well established that ABA levels in leaves and guard cells can increase following the imposition of turgor loss or water stress (Pierce and Raschke, 1980; Harris et al., 1988; Harris and Outlaw, 1991). However, only a few studies have reported increases in foliar ABA level in response to high VPD (Bauerle et al., 2004; Giday et al., 2013), and none have investigated whether these observed dynamic changes or differences in ABA level were functionally relevant for stomatal control. In addition, no study has quantified the levels of ABA in guard cells during a transition in VPD.Here, we investigate the relative importance of ABA for the stomatal response to VPD in whole plants, sampled from across the vascular land plant lineage. We provide, to our knowledge, the first functional assessment of changes in ABA levels driving stomatal responses to VPD as well as critically investigate the recent suggestion that stomatal responses to VPD are driven by an autonomous guard cell synthesis of ABA.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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