首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
2.
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   

3.
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   

4.
The effect of nitrogen (N) stress on the pool system supplying currently assimilated and (re)mobilized N for leaf growth of a grass was explored by dynamic 15N labeling, assessment of total and labeled N import into leaf growth zones, and compartmental analysis of the label import data. Perennial ryegrass (Lolium perenne) plants, grown with low or high levels of N fertilization, were labeled with 15NO3/14NO3 from 2 h to more than 20 d. In both treatments, the tracer time course in N imported into the growth zones fitted a two-pool model (r2 > 0.99). This consisted of a “substrate pool,” which received N from current uptake and supplied the growth zone, and a recycling/mobilizing “store,” which exchanged with the substrate pool. N deficiency halved the leaf elongation rate, decreased N import into the growth zone, lengthened the delay between tracer uptake and its arrival in the growth zone (2.2 h versus 0.9 h), slowed the turnover of the substrate pool (half-life of 3.2 h versus 0.6 h), and increased its size (12.4 μg versus 5.9 μg). The store contained the equivalent of approximately 10 times (low N) and approximately five times (high N) the total daily N import into the growth zone. Its turnover agreed with that of protein turnover. Remarkably, the relative contribution of mobilization to leaf growth was large and similar (approximately 45%) in both treatments. We conclude that turnover and size of the substrate pool are related to the sink strength of the growth zone, whereas the contribution of the store is influenced by partitioning between sinks.This article examines the nitrogen (N) supply system of growing grass leaves, and it investigates how functional and kinetic properties of this system are affected by N stress. The N supply of growing leaves is a dominant target of whole-plant N metabolism. This is primarily related to the high N demand of the photosynthetic apparatus and the related metabolic machinery of new leaves (Evans, 1989; Makino and Osmond, 1991; Grindlay, 1997; Lemaire, 1997; Wright et al., 2004; Johnson et al., 2010; Maire et al., 2012). The N supply system, as defined here, is an integral part of the whole plant: it includes all N compounds that supply leaf growth. Hence, it integrates all events between the uptake of N from the environment (source), intermediate uses in other processes of plant N metabolism, and the eventual delivery to the leaf growth zone (sink; Fig. 1). N that does not ultimately serve leaf growth is not included in this system; all N that serves leaf growth is included, irrespective of its localization in the plant. Conceptually, two distinct sources supply N for leaf growth: N from current uptake and assimilation that is directly transferred to the growing leaf (“directly transferred N”) and N from turnover/redistribution of organic compounds (“mobilized N”).Open in a separate windowFigure 1.Schematic representation of N fluxes in the leaf growth zone and in the N supply system of leaf growth in a grass plant. A, Scheme of a growing leaf, with its growth zone (including zones of cell division, expansion, and maturation) and recently produced tissue (RPT). N import (I; μg h−1) into the growth zone is mostly in the form of amino acids. Inside the growth zone, the nitrogenous substrate is used in new tissue construction. Then, N export (E; μg h−1) is in the form of newly formed, fully expanded nitrogenous tissue (tissue-bound export with RPT) and is calculated as leaf elongation rate (LER; mm h−1) times the lineal density of N in RPT (ρ; μg mm−1): E = LER × ρ (Lattanzi et al., 2004). In a physiological steady state, import equals export (I = E) and the N content of the growth zone (G; μg [not shown]) is constant. Labeled N import into the growth zone (Ilab) commences shortly after labeling of the nutrient solution with 15N. The labeled N content of the growth zone (Glab; μg) increases over time (dGlab/dt) until it eventually reaches isotopic saturation (Fig. 2B). Similarly, the lineal density of labeled N in RPTlab) increases until it approaches ρ. At any time, the export of labeled N in RPT (Elab) equals the concurrent ρlab × LER. The import of labeled N is obtained as Ilab = Elab + dGlab/dt (Lattanzi et al., 2005) and considers the increasing label content in the growth zone during labeling. The fraction of labeled N in the import flux (flab I) is calculated as flab I = Ilab/I. The time course of flab I (Fig. 3) reflects the kinetic properties of the N supply system of leaf growth (C). B, Scheme of a vegetative grass plant (reduced to a rooted tiller with three leaves) with leaf growth zone. N import into the growth zone (I) originates from (1) N taken up from the nutrient solution that is transferred directly to the growth zone following assimilation (directly transferred N) and (2) N derived from turnover/redistribution of stores (mobilized N). The store potentially includes proteins in all mature and senescing tissue in the shoot and root of the entire plant. As xylem, phloem, and associated transfer cells/tissue provide for a vascular network that connects all parts of the plant, the mobilized N may principally originate from any plant tissue that exhibits N turnover/mobilization. The fraction of total N uptake that is allocated to the N supply system of the growth zone equals U (see model in C). The fraction of total mobilized N allocated to the growth zone equals M (see model in C). C, Compartmental model of the source-sink system supplying N to the leaf growth zone, as shown by Lattanzi et al. (2005) and used here. Newly absorbed N (U; μg h−1) enters a substrate pool (Q1); from there, the N is either imported directly into the growth zone (I) or exchanged with a store (Q2). Q1 integrates the steps of transport and assimilation that precede the translocation to the growth zone. Q2 includes all proteins that supply N for leaf growth during their turnover and mobilization. The parameters of the model, including the (relative) size and turnover of pools Q1 and Q2, the deposition into the store (D; μg h−1), and the mobilization from the store (M; μg h−1), and the contribution of direct transfer relative to mobilization to the N supply of the growth zone are obtained by fitting the compartmental model to the flab I data (A) obtained in dynamic 15N labeling experiments (for details, see “Materials and Methods”). During physiological steady state, the sizes of Q1 and Q2 are constant, I = U, and M = D. [See online article for color version of this figure.]Amino acids are the predominant form in which N is supplied for leaf growth in grasses, and incorporation in new leaf tissue occurs mainly in the leaf growth zone (Gastal and Nelson, 1994; Amiard et al., 2004). This is a heterotrophic piece of tissue that includes the zones of cell division and elongation, is located at the base of the leaf, and is encircled by the sheath of the next older leaf (Volenec and Nelson, 1981; MacAdam et al., 1989; Schnyder et al., 1990; Kavanová et al., 2008). As most N is taken up in the form of nitrate but supplied to the growth zone in the form of amino acids, the path of directly transferred N includes a series of metabolic and transport steps. These include transfer to and loading into the xylem, xylem transport and unloading, reduction and ammonium assimilation, cycling through photorespiratory N pools, amino acid synthesis, loading into the phloem, and transport to the growth zone (Hirel and Lea, 2001; Novitskaya et al., 2002; Stitt et al., 2002; Lalonde et al., 2003; Dechorgnat et al., 2011). The time taken to pass through this sequence is unknown at present, as is the effect of N deficiency on that time. Also, it is not known how much N is contained in, and moving through, the different compartments that supply leaf growth with currently assimilated N.At the level of mature organs, mainly leaves, there is considerable knowledge about N turnover and redistribution. Much less is known about the fate of the mobilized N and its actual use in sink tissues like the leaf growth zone. The processes in mature organs are associated with the maintenance metabolism of proteins, organ senescence, and adjustments in leaf protein levels to decreasing irradiance inside growing canopies when leaves become shaded by overtopping newer ones (Evans, 1993; Vierstra, 1993; Hikosaka et al., 1994; Anten et al., 1995; Hirel et al., 2007; Jansson and Thomas, 2008; Moreau et al., 2012). N mobilization in shaded leaves supports the optimization of photosynthetic N use efficiency at plant and canopy scale (Field, 1983; Evans, 1993; Anten et al., 1995), it reduces the respiratory burden of protein maintenance costs (Dewar et al., 1998; Amthor, 2000; Cannell and Thornley, 2000), and it provides a mechanism for the conservation of the most frequently growth-limiting nutrient (Aerts, 1996). Mobilization of N involves protein turnover and net degradation (Huffaker and Peterson, 1974), redistribution in the form of amino acids (Simpson and Dalling, 1981; Simpson et al., 1983; Hörtensteiner and Feller, 2002), and (at least) some of the mobilized N is supplied to new leaf growth (Lattanzi et al., 2005).N fertilizer supply has multiple direct and indirect effects on plant N metabolism (Stitt et al., 2002; Schlüter et al., 2012). In particular, it modifies the N content of newly produced leaves, leaf longevity/senescence, and the dynamics of light distribution inside expanding canopies (Evans, 1983, 1989; Lötscher et al., 2003; Moreau et al., 2012). Thus, N fertilization influences the availability of recyclable N. At the same time, it augments the availability of directly transferable N to leaf growth. The net effect of these factors on the importance of mobilized versus directly transferred N substrate for leaf growth is not known. Also, it is unknown how N fertilization influences the functional characteristics of the N supply system, such as the size and turnover of its component pools.The assessment of the importance of directly transferred versus mobilized N for leaf growth requires studies at the sink end of the system (i.e. investigations of the N import flux into the leaf growth zone). Directly transferred N and mobilized N can be distinguished on the basis of their residence time in the plant, the time between uptake from the environment and import into the leaf growth zone: direct transfer involves a short residence time (fast transfer), whereas mobilized N resides much longer in the plant before it is delivered to the growth zone (slow transfer; De Visser et al., 1997; Lattanzi et al., 2005). Such studies require dynamic labeling of the N taken up by the plant (Schnyder and de Visser, 1999) and monitoring of the rate and isotopic composition/label content of N import into the leaf growth zone (Lattanzi et al., 2005). For grass plants in a physiological steady state, N import and the isotopic composition of the imported N are calculated from the leaf elongation rate and the lineal density of N in newly formed tissue (Fig. 1A; Lattanzi et al., 2004) and the change of tracer content in the leaf growth zone and recently produced leaf tissue over time (Lattanzi et al., 2005). Such data reveal the temporal change of the fraction of labeled N in the N import flux (flab I), which then can be used to characterize the N supply system of leaf growth via compartmental modeling. So far, there is only one study that has partially characterized this system (Lattanzi et al., 2005): this work was conducted with a C3 grass, perennial ryegrass (Lolium perenne), and a C4 grass, Paspalum dilatatum, growing in mixed stands and indicated that two interconnected N pools supplied the leaf growth zone in both species: a “substrate pool” (Q1), which provided a direct route for newly absorbed and assimilated N import into the leaf growth zone (directly transferred N), and a mobilizing “store” (Q2), which supplied N to the leaf growth zone via the substrate pool (Fig. 1C). The relative contribution of mobilization from the store was least important in the fast-growing, dominant individuals and most important in subordinate, shaded individuals. That work did not address the role of N deficiency, and the limited short-term resolution of the study (labeling intervals of 24 h or greater) precluded an analysis of the fast-moving parts of the system.Accordingly, this work addresses the following questions. How does N deficiency influence the substrate supply system of the leaf growth sink in terms of the number, size, and turnover (half-life) of its kinetically distinct pools? How does N deficiency affect the relationship between directly transferred and mobilized N for leaf growth? And what additional insight on the compartmental structure of the supply system is obtained when the short-term resolution of the analysis is increased by 1 order of magnitude? The work was performed with vegetative plants of perennial ryegrass grown in constant conditions with either a low (1.0 mm; termed low N) or high (7.5 mm; high N) nitrate concentration in the nutrient solution. In both treatments, a large number of plants were dynamically labeled with 15N over a wide range of time intervals (2 h to more than 20 d). The import of total N and 15N tracer into growth zones was estimated at the end of each labeling interval. Tracer data were analyzed with compartmental models following principles detailed by Lattanzi et al. (2005, 2012) and Lehmeier et al. (2008) to address the specific questions. Previous articles reported on root and shoot respiration (Lehmeier et al., 2010) and cell division and expansion in leaf growth zones (Kavanová et al., 2008) in the same experiment.  相似文献   

5.
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   

6.
Freeze-thaw events can affect plant hydraulics by inducing embolism. This study analyzed the effect of temperature during the freezing process on hydraulic conductivity and ultrasonic emissions (UE). Stems of 10 angiosperms were dehydrated to a water potential at 12% percentage loss of hydraulic conductivity (PLC) and exposed to freeze-thaw cycles. The minimal temperature of the frost cycle correlated positively with induced PLC, whereby species with wider conduits (hydraulic diameter) showed higher freeze-thaw-induced PLC. Ultrasonic activity started with the onset of freezing and increased with decreasing subzero temperatures, whereas no UE were recorded during thawing. The temperature at which 50% of UE were reached varied between −9.1°C and −31.0°C across species. These findings indicate that temperatures during freezing are of relevance for bubble formation and air seeding. We suggest that species-specific cavitation thresholds are reached during freezing due to the temperature-dependent decrease of water potential in the ice, while bubble expansion and the resulting PLC occur during thawing. UE analysis can be used to monitor the cavitation process and estimate freeze-thaw-induced PLC.Xylem embolism is a limiting factor for plant survival and distribution (Choat et al., 2012; Charrier et al., 2013). Two major factors can induce embolism in the xylem of plants: drought and freeze stress. Freeze-thaw-induced embolism is caused by bubbles formed during freezing that then expand on thawing (Lemoine et al., 1999; Hacke and Sperry, 2001; Cruiziat et al., 2002; Tyree and Zimmermann, 2002). As wider conduits contain more gas and form larger bubbles, which expand at less negative tension, conduit diameter and xylem sap tension are critical for the formation of freeze-thaw-induced embolism (Davis et al., 1999; Pittermann and Sperry, 2003). Accordingly, Mayr and Sperry (2010) observed a loss of conductivity only when samples were under critical tension during thawing. Under drought stress, tension in the xylem sap increases the sensitivity to embolism generated by successive freeze-thaw cycles (Mayr et al., 2003, 2007).Ultrasonic emissions (UE) analysis can be used to detect cavitation events in wood. It is unclear how well related UE are to cavitation events, as they are extracted from continuous acoustic emissions and depend on set definitions. However, UE analysis has been proven effective for monitoring drought-induced embolism in the laboratory (Pena and Grace, 1986; Salleo and Lo Gullo, 1986; Borghetti et al., 1993; Salleo et al., 2000) as well as in field experiments (Ikeda and Ohtsu, 1992; Jackson et al., 1995; Jackson and Grace, 1996; Hölttä et al., 2005; Ogaya and Penuelas, 2007). In a cavitating conduit, signals are probably produced by the disruption of the water column and subsequent tension relaxation of cell walls.UE have also been detected during freezing events, but the origin of these signals was less clear. In some cases, UE were observed during thawing, which are thus probably related to embolism formation according to the classic thaw-expansion hypothesis (Mayr and Sperry, 2010); however, all species studied have produced UE on freezing, which cannot yet be explained (Raschi et al., 1989; Kikuta and Richter, 2003; Mayr et al., 2007; Mayr and Sperry, 2010; Mayr and Zublasing, 2010). The low solubility of gases in ice prompted the idea that air bubbles expulsed from the ice structure produce UE near the ice-liquid interface (Sevanto et al., 2012). As the water potential of ice is strongly temperature dependent, the minimum temperature during freezing might be a relevant factor. Numerous studies have analyzed UE patterns during freeze-thaw cycles in conifers (Mayr et al., 2007; Mayr and Sperry, 2010; Mayr and Zublasing, 2010) or angiosperms (Weiser and Wallner, 1988; Kikuta and Richter, 2003), but few of them measured embolism concomitantly. Percentage loss of hydraulic conductivity (PLC) was only measured in a few studies and only in conifers (Mayr et al., 2007; Mayr and Sperry, 2010).In this study, we analyzed the effect of freeze-thaw cycles on the hydraulic conductivity and ultrasonic activity in 10 angiosperm species. We hypothesized that (1) the extent of freeze-thaw-induced embolism depends on xylem anatomy (related to conduit diameter) and minimal temperature (related to the water potential of ice); (2) ultrasonic activity is also influenced by anatomy and temperature; and (3) PLC and UE are positively correlated. PLC was measured in 10 angiosperm species after freeze-thaw cycles at different minimal temperatures (−10 to −40°C). Furthermore, UE were recorded during a freeze-thaw cycle down to −40°C.  相似文献   

7.
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   

8.
Red algae represent an evolutionarily important group that gave rise to the whole red clade of photosynthetic organisms. They contain a unique combination of light-harvesting systems represented by a membrane-bound antenna and by phycobilisomes situated on thylakoid membrane surfaces. So far, very little has been revealed about the mobility of their phycobilisomes and the regulation of their light-harvesting system in general. Therefore, we carried out a detailed analysis of phycobilisome dynamics in several red alga strains and compared these results with the presence (or absence) of photoprotective mechanisms. Our data conclusively prove phycobilisome mobility in two model mesophilic red alga strains, Porphyridium cruentum and Rhodella violacea. In contrast, there was almost no phycobilisome mobility in the thermophilic red alga Cyanidium caldarium that was not caused by a decrease in lipid desaturation in this extremophile. Experimental data attributed this immobility to the strong phycobilisome-photosystem interaction that highly restricted phycobilisome movement. Variations in phycobilisome mobility reflect the different ways in which light-harvesting antennae can be regulated in mesophilic and thermophilic red algae. Fluorescence changes attributed in cyanobacteria to state transitions were observed only in mesophilic P. cruentum with mobile phycobilisomes, and they were absent in the extremophilic C. caldarium with immobile phycobilisomes. We suggest that state transitions have an important regulatory function in mesophilic red algae; however, in thermophilic red algae, this process is replaced by nonphotochemical quenching.Photosynthetic light reactions are mediated by pigment-binding protein complexes located either inside the thylakoid membrane (e.g. chlorophyll-binding proteins of both photosystems) or associated on the membrane surface (e.g. phycobilisomes [PBsomes] in cyanobacteria and red algae). Recent progress in structural biology has allowed the construction of high-resolution structural models of most photosynthetic protein complexes (for review, see Fromme, 2008) together with their large-scale organization into supercomplexes (for review, see Dekker and Boekema, 2005). However, the dynamics of these supercomplexes and the mobility of particular light-harvesting proteins in vivo are still poorly understood (for review, see Mullineaux, 2008a; Kaňa, 2013; Kirchhoff, 2014) The importance of protein mobility in various photosynthetic processes, like nonphotochemical quenching and state transitions, has been explored mostly based on indirect in vitro experiments, including single-particle analysis (Kouřil et al., 2005), or by biochemical methods (Betterle et al., 2009; Caffarri et al., 2009). Recent studies on the mobility of light-harvesting proteins using live-cell imaging (for review, see Mullineaux, 2008a; Kaňa, 2013) have elucidated the importance of protein mobility for photosynthetic function (Joshua and Mullineaux, 2004; Joshua et al., 2005; Goral et al., 2010, 2012; Johnson et al., 2011). In addition, the redistribution of respiratory complexes in cyanobacterial thylakoid membranes plays an essential role in controlling electron flow (Liu et al., 2012).It is generally accepted that the mobility of most of the transmembrane photosynthetic proteins is very restricted in the thylakoid. The typical effective diffusion coefficient of photosynthetic proteins is somewhere between 0.01 and 0.001 μm−2 s−1 (Kaňa, 2013). A similar restriction in membrane protein mobility has also been described for bacterial membranes (Dix and Verkman, 2008; Mika and Poolman, 2011). In fact, this is very different in comparison with what we know for other eukaryotic membranes (e.g. plasma membrane and endoplasmic reticulum), where membrane-protein diffusion can be faster by 1 or 2 orders of magnitude (Lippincott-Schwartz et al., 2001). Therefore, macromolecular crowding of proteins has been used to rationalize the restricted protein mobility in thylakoid membranes of chloroplasts (Kirchhoff, 2008a, 2008b). Indeed, atomic force microscopy studies have shown that there is a dense packing and interaction of complexes in the photosynthetic membranes (Liu et al., 2011). Therefore, the diffusion of photosynthetic proteins in the thylakoid membrane is rather slow, and it increases only in less crowded parts of thylakoids (Kirchhoff et al., 2013). The current model of photosynthetic protein mobility thus proposes the immobility of protein supercomplexes, such as PSII (Mullineaux et al., 1997; Kirchhoff, 2008b), with only a small mobile fraction of chlorophyll-binding proteins represented by external antennae of photosystems, including light harvesting complex of PSII in higher plants (Consoli et al., 2005; Kirchhoff et al., 2008) or iron stress-induced chlorophyll-binding protein A in cyanobacteria (Sarcina and Mullineaux, 2004).The restricted mobility of internal membrane supercomplexes (photosystems) contrasts with the relatively mobile PBsomes (Mullineaux et al., 1997; Sarcina et al., 2001). PBsomes are sizeable biliprotein supercomplexes (5–10 MD) attached to the thylakoid membrane surface with dimensions of approximately 64 × 42 × 28 nm (length × width × height; Arteni et al., 2008; Liu et al., 2008a). PBsomes are composed of chromophore-bearing phycobiliproteins and colorless linker polypeptides (Adir, 2005; Liu et al., 2005). They serve as the main light-harvesting antennae in various species, including cyanobacteria, red algae, glaucocystophytes, and cryptophytes. Although a single PBsome is composed of hundreds of biliproteins, absorbed light energy is efficiently transferred toward a specific biliprotein that functions as a terminal energy emitter (Glazer, 1989). From there, energy can be transferred to either PSI or PSII and used in photosynthesis (Mullineaux et al., 1990; Mullineaux, 1992, 1994). In typical prokaryotic cyanobacteria and eukaryotic red algae, PBsomes are composed of two main parts: (1) allophycocyanin (APC) core proteins adjacent to the thylakoid membrane; and (2) peripheral rod proteins made from phycocyanin only or from a combination of phycocyanin together with phycoerythrin. Such complex and modular composition allows for different spectroscopic properties of PBsomes and thus their complementary absorption in the spectral region that is not covered by chlorophyll-binding proteins.PBsome mobility has been studied only in a few types of cyanobacteria (for review, see Kaňa, 2013). PBsomes have been recognized as a mobile element with an effective diffusion coefficient of about 0.03 μm2 s−1 for Synechococcus sp. PCC 7942 (Mullineaux et al., 1997; Sarcina et al., 2001). The effective diffusion coefficient value depends on lipid composition, temperature, and the size of the PBsome (Sarcina et al., 2001). The diffusion coefficient reflects PBsome mobility, but it is not affected singularly by physical diffusion processes, and the role of PBsome-photosystem interaction is an open question (Kaňa, 2013). PBsome mobility seems to be related to the requirement of light-induced PBsome redistribution during state transitions (Joshua and Mullineaux, 2004). The mechanism of state transitions in cyanobacteria is still rather questionable (for review, see Kirilovsky et al., 2014). As PSII seems to be immobile, it has been suggested that PBsomes interact with photosystems only transiently and that physical redistribution (diffusion) of PBsomes is crucial for the state transition (Mullineaux et al., 1997). The importance of such long-distance diffusions, however, should be tested experimentally in more detail (Kaňa, 2013), as an alternative theory of the state transition proposed only slight PBsome movement (shifting) between photosystems (McConnell et al., 2002). However, in both cases, PBsome mobility (i.e. the PBsome’s ability to move) is required (Kaňa, 2013).Red algae are the eukaryotic representatives of phototrophs containing PBsomes (Su et al., 2010). They represent the ancestor of photosynthetic microorganisms from the red clade of photosynthesis (Yoon et al., 2006; Wang et al., 2013), which includes various model organisms such as diatoms, chromerids, or dinoflagellates. Red algae contain a unique combination of antennae systems on their membrane surfaces, which are formed mostly by hemispherical PBsomes (Mimuro and Kikuchi, 2003; Arteni et al., 2008). Red algae also contain transmembrane light-harvesting antennae (Vanselow et al., 2009; Neilson and Durnford, 2010; Green, 2011) associated mostly with PSI (Wolfe et al., 1994). Therefore, red algae represent a functionally important eukaryotic model organism; however, few facts are known about the regulation of its light-harvesting efficiency, although it seems to be connected with photoprotection in the reaction center (Delphin et al., 1996, 1998; Krupnik et al., 2013). The presence of photoprotective NPQ in PBsomes of prokaryotic cyanobacteria has been conclusively proven (Kirilovsky et al., 2014); however, this mechanism seems to be missing in eukaryotic phycobiliproteins of cryptophytes (Kaňa et al., 2012b) and red algae. Moreover, the presence (or absence) of PBsome mobility has not been confirmed conclusively (Liu et al., 2009).Therefore, we carried out a detailed study of PBsome mobility in red algal chloroplasts to determine the role of mobility in the regulation of light-harvesting efficiency. We found that red alga PBsomes are a mobile protein complex with effective diffusion coefficient between 2.7 × 10−3 and 13 × 10−3 μm−2 s−1 in all studied mesophilic strains. It contrasted with PBsomes in extremophilic red algal strains (Cyanidium caldarium), where PBsome mobility under physiological conditions was highly restricted (effective diffusion coefficient of approximately 0.6 × 10−3 μm−2 s−1). The restriction of PBsome mobility in extremophilic C. caldarium was due to a tight interaction of PBsomes with both photosystems and not to changes in lipid desaturation, an effect typical for extremophiles. The PBsome-photosystem interaction was weakened for C. caldarium grown at suboptimal temperatures, resulting in a pronounced increase in PBsome mobility thanks to PBsome decoupling from the photosystem. This result shows that PBsome mobility in this strain is limited by the strength of the PBsome-photosystem interaction rather than by the restriction of diffusion by factors such as macromolecular crowding. Moreover, our study allows us to describe two different models of light-harvesting antenna regulation in red algae. In mesophilic strains (Porphyridium cruentum and Rhodella violacea), absorbed light is redistributed between photosystems in a process of state transition that requires PBsome mobility. On the contrary, in extremophilic C. caldarium, PBsome are strongly coupled to photosystems and excess light is dissipated by a process of nonphotochemical quenching, as has been described recently (Krupnik et al., 2013).  相似文献   

9.
Enhancing nitrogen use efficiency (NUE) in crop plants is an important breeding target to reduce excessive use of chemical fertilizers, with substantial benefits to farmers and the environment. In Arabidopsis (Arabidopsis thaliana), allocation of more NO3 to shoots was associated with higher NUE; however, the commonality of this process across plant species have not been sufficiently studied. Two Brassica napus genotypes were identified with high and low NUE. We found that activities of V-ATPase and V-PPase, the two tonoplast proton-pumps, were significantly lower in roots of the high-NUE genotype (Xiangyou15) than in the low-NUE genotype (814); and consequently, less vacuolar NO3 was retained in roots of Xiangyou15. Moreover, NO3 concentration in xylem sap, [15N] shoot:root (S:R) and [NO3] S:R ratios were significantly higher in Xiangyou15. BnNRT1.5 expression was higher in roots of Xiangyou15 compared with 814, while BnNRT1.8 expression was lower. In both B. napus treated with proton pump inhibitors or Arabidopsis mutants impaired in proton pump activity, vacuolar sequestration capacity (VSC) of NO3 in roots substantially decreased. Expression of NRT1.5 was up-regulated, but NRT1.8 was down-regulated, driving greater NO3 long-distance transport from roots to shoots. NUE in Arabidopsis mutants impaired in proton pumps was also significantly higher than in the wild type col-0. Taken together, these data suggest that decrease in VSC of NO3 in roots will enhance transport to shoot and essentially contribute to higher NUE by promoting NO3 allocation to aerial parts, likely through coordinated regulation of NRT1.5 and NRT1.8.China is the largest consumer of nitrogen (N) fertilizer in the world; however, the average N use efficiency (NUE) in fertilizer is only around 35%, suggesting considerable potential for improvements (Shen et al., 2003; Wang et al., 2014). With the high amounts of N-fertilizer being used, crop yields are declining in some areas, where application is exceeding the optimum required for local field crops (Shen et al., 2003; Miller and Smith, 2008; Xu et al., 2012). The extremely low NUE results in waste of resources and environmental contamination, and also presents serious hazards for human health (Xu et al., 2012; Chen et al., 2014). Consequently, exploiting the maximum potential for improving NUE in crop plants will have practical significance for agriculture production and the environment (Zhang et al., 2010; Schroeder et al., 2013; Wang et al., 2014). Elucidating the genetic and physiological regulatory mechanisms governing NUE in plants will allow breeding crops and varieties with higher NUE.Ammonium (NH4+) and nitrate (NO3) are the main N species absorbed and utilized by crops, and NO3 accumulation and utilization are of major emphasis for N nutrient studies in dry land crops, such as Brassica napus. Several studies revealed the close relationship between NO3 content and NUE in plant tissues (Shen et al., 2003; Zhang et al., 2012; Tang et al., 2013; Han et al., 2015a). When plants are sufficiently illuminated, NO3 assimilation efficiency significantly increase in shoots compared with roots (Smirnoff and Stewart, 1985; Tang et al., 2013). Consequently, under daytime with optimal illumination, higher proportion of NO3 in plant tissue is transported from root to shoot, as an advantageous physiological adaptation that reduces the cost of energy for metabolism (Tang et al., 2013). NO3 assimilation in plant shoots can therefore take advantage of solar energy while improving NUE (Smirnoff and Stewart, 1985; Andrews, 1986; Tang et al., 2012, 2013).The NO3 long-distance transport and distribution between root and shoot is regulated by two genes encoding long transport mechanisms. NRT1.5 is responsible for xylem NO3 loading, while NRT1.8 is responsible for xylem NO3 unloading (Lin et al., 2008; Li et al., 2010). Expression of the two genes is influenced by NO3 concentration. NRT1.5 is strongly induced by NO3 (Lin et al., 2008), while NRT1.8 expression is extremely up-regulated in nrt1.5 mutants (Chen et al., 2012). A negative correlation between the extents of expression of the two genes was observed when plants are subjected to abiotic stresses (Chen et al., 2012). Moreover, expression of NRT1.5 is strongly inhibited by 1-aminocyclopropane-1-carboxylic acid (ACC) and methyl jasmonate (MeJA), whereas the expression of NRT1.8 is significantly up-regulated (Zhang et al., 2014). Based on these studies, we argue that the expression and functioning of NO3 long-distance transport genes NRT1.5 and NRT1.8 are regulated by cytosolic NO3 concentration. In addition, the vacuolar and cytosolic NO3 distribution is likely regulated by proton pumps located within the tonoplast (V-ATPase and V-PPase; Granstedt and Huffaker, 1982; Glass et al., 2002; Krebs et al., 2010). Therefore, NO3 use efficiency must be affected by NO3 long-distant transport (between shoot and root) and short-distant transport (between vacuole and cytosol). However, the physiological mechanisms controlling this regulation are still obscure.Previous studies showed that the chloride channel protein (CLCa) is mainly responsible for vacuole NO3 short-distance transport, as it is the main channel for NO3 movement between the vacuoles and cytosol (De Angeli et al., 2006; Wege et al., 2014). The vacuole proton-pumps (V-ATPase and V-PPase) located in the tonoplast supply energy for active transport of NO3 and accumulation within the vacuole (Gaxiola et al., 2001; Brüx et al., 2008; Krebs et al., 2010). Despite the fact about 90% of the volume of mature plant cells is occupied by vacuoles, vacuolar NO3 cannot be efficiently assimilated because the enzyme nitrate reductase (NR) is cytosolic (Shen et al., 2003; Han et al., 2015a). However, retranslocation of NO3 from the vacuole to the cytosol will permit its immediate assimilation and utilization.Generally, NO3 concentrations in plant cell vacuoles and the cytoplasm are in the range of 30–50 mol m−3 and 3–5 mol m−3, respectively (Martinoia et al., 1981, 2000). Because vacuoles are obviously the organelle for high NO3 accumulation and storage in plant tissues, their function in NO3 use efficiency cannot be ignored (Martinoia et al., 1981; Zhang et al., 2012; Han et al., 2015b). NO3 assimilatory system in the cytoplasm is sufficient for its assimilation when it is transported out of the vacuoles. Therefore, NO3 use efficiency could in part be dependent on vacuolar-cytosolic NO3 short-distance transport in plant tissues (Martinoia et al., 1981; Shen et al., 2003; Zhang et al., 2012; Han et al., 2015a).Evidently, NO3 use efficiency is regulated by both NO3 long-distance transport from root to shoot and short-distance transport and distribution between vacuoles and cytoplasm within cells (Glass et al., 2002; Dechorgnat et al., 2011; Han et al., 2015a). Although vacuoles compartment excess NO3 that accumulates in plant cells (Granstedt and Huffaker, 1982; Krebs et al., 2010), neither NO3 inducible NR genes (NIA1 and NIA2; Fan et al., 2007; Han et al., 2015a) nor the NO3 long-distance transport gene NRT1.5 (Lin et al., 2008) are regulated by vacuolar NO3, even though they are essential for NO3 assimilation. Only NO3 transported from the vacuole to the cytosol can play a role in regulating NO3 inducible genes. Consequently, we argue that both NO3 assimilation in cells and its long-distance transport from root to shoot are regulated by cytosolic NO3 concentration. However, this hypothesis needs to be substantiated. The mechanisms underlying both NO3 short-distance (Gaxiola et al., 2001; De Angeli et al., 2006; Brüx et al., 2008; Krebs et al., 2010) and long-distance transport (Lin et al., 2008; Li et al., 2010) have been previously investigated, yet the underlying mechanisms regulating the flux of NO3 and the obvious relationship between the two transport pathways, as well as their relation to NUE, are not well understood.The NRT family of genes play a partial role in vacuolar NO3 accumulation in petioles (Chiu et al., 2004) and seed tissues (Chopin et al., 2007), whereas the proton pumps and CLCa system in the tonoplast play a major role in accumulating NO3 in vacuoles (Gaxiola et al., 2001; De Angeli et al., 2006; Brüx et al., 2008; Krebs et al., 2010). The vacuolar NO3 short-distance transport system is spread throughout the plant tissues and is the principal means by which vacuolar NO3 short-distance transport and distribution is controlled (De Angeli et al., 2006; Krebs et al., 2010).The NRT genes seem to work synergistically to control NO3 long-distance transport between roots and shoots. NRT1.9 is responsible for NO3 loading into the phloem (Wang and Tsay, 2011), whereas NO3 loading and unloading into xylem are regulated by NRT1.5 and NRT1.8, respectively (Lin et al., 2008; Li et al.; 2010). Phloem transport mainly involves organic N; the inorganic-N (NO3) concentrations in the phloem sap are typically very low, ranging from one-tenth to one-hundredth of that of the inorganic-N in xylem sap (Lin et al., 2008; Fan et al., 2009). Therefore, this study focused on NO3 short-distance transport mediated through the tonoplast proton pumps and the CLCa system and the long-distant transport mechanisms responsible for xylem NO3 loading and unloading via NRT1.5 and NRT1.8, respectively.Questions related to how long- and short-distance transport of NO3 are coupled in plant tissues and their role in determining NUE were addressed using a pair of high- and low-NUE B. napus genotypes and Arabidopsis (Arabidopsis thaliana). Application of proton pump inhibitors and ACC in the former, and use of mutants with defective proton pumps in the latter, allowed experimental distinction of the physiological mechanisms regulating these processes. Data presented here provide strong evidence from both model plants supporting this linkage and strongly suggest that cytosolic NO3 concentration in roots regulates NO3 long-distance transport from roots to shoots. We also investigated how NO3 concentration in plant tissues would be affected by NO3 long-distance transport, vacuolar NO3 sequestration, and the ensuing relationship with NO3 use efficiency. We also proposed the physiological mechanisms likely to be important for enhancing NO3 use efficiency in plants. These findings will provide scientific rationales for improving NUE in important industrial and food crops.  相似文献   

10.
11.
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   

12.
In plant cells, secretory and endocytic routes intersect at the trans-Golgi network (TGN)/early endosome (EE), where cargos are further sorted correctly and in a timely manner. Cargo sorting is essential for plant survival and therefore necessitates complex molecular machinery. Adaptor proteins (APs) play key roles in this process by recruiting coat proteins and selecting cargos for different vesicle carriers. The µ1 subunit of AP-1 in Arabidopsis (Arabidopsis thaliana) was recently identified at the TGN/EE and shown to be essential for cytokinesis. However, little was known about other cellular activities affected by mutations in AP-1 or the developmental consequences of such mutations. We report here that HAPLESS13 (HAP13), the Arabidopsis µ1 adaptin, is essential for protein sorting at the TGN/EE. Functional loss of HAP13 displayed pleiotropic developmental defects, some of which were suggestive of disrupted auxin signaling. Consistent with this, the asymmetric localization of PIN-FORMED2 (PIN2), an auxin transporter, was compromised in the mutant. In addition, cell morphogenesis was disrupted. We further demonstrate that HAP13 is critical for brefeldin A-sensitive but wortmannin-insensitive post-Golgi trafficking. Our results show that HAP13 is a key link in the sophisticated trafficking network in plant cells.Plant cells contain sophisticated endomembrane compartments, including the endoplasmic reticulum, the Golgi, the trans-Golgi network (TGN)/early endosome (EE), the prevacuolar compartments/multivesicular bodies (PVC/MVB), various types of vesicles, and the plasma membrane (PM; Ebine and Ueda, 2009; Richter et al., 2009). Intracellular protein sorting between the various locations in the endomembrane system occurs in both secretory and endocytic routes (Richter et al., 2009; De Marcos Lousa et al., 2012). Vesicles in the secretory route start at the endoplasmic reticulum, passing through the Golgi before reaching the TGN/EE, while vesicles in the endocytic route start from the PM before reaching the TGN/EE (Dhonukshe et al., 2007; Viotti et al., 2010). The TGN/EE in Arabidopsis (Arabidopsis thaliana) is an independent and highly dynamic organelle transiently associated with the Golgi (Dettmer et al., 2006; Lam et al., 2007; Viotti et al., 2010), distinct from the animal TGN. Once reaching the TGN/EE, proteins delivered by their vesicle carriers are subject to further sorting, being incorporated either into vesicles that pass through the PVC/MVB before reaching the vacuole for degradation or into vesicles that enter the secretory pathway for delivery to the PM (Ebine and Ueda, 2009; Richter et al., 2009). Therefore, the TGN/EE is a critical sorting compartment that lies at the intersection of the secretory and endocytic routes.Fine-tuned control of intracellular protein sorting at the TGN/EE is essential for plant development (Geldner et al., 2003; Dhonukshe et al., 2007, 2008; Richter et al., 2007; Kitakura et al., 2011; Wang et al., 2013). An auxin gradient is crucial for pattern formation in plants, whose dynamic maintenance requires the polar localization of auxin efflux carrier PINs through endocytic recycling (Geldner et al., 2003; Blilou et al., 2005; Paciorek et al., 2005; Abas et al., 2006; Jaillais et al., 2006; Dhonukshe et al., 2007; Kleine-Vehn et al., 2008). Receptor-like kinases (RLKs) have also been recognized as major cargos undergoing endocytic trafficking, which are either recycled back to the PM or sent for vacuolar degradation (Geldner and Robatzek, 2008; Irani and Russinova, 2009). RLKs are involved in most if not all developmental processes of plants (De Smet et al., 2009).Intracellular protein sorting relies on sorting signals within cargo proteins and on the molecular machinery that recognizes sorting signals (Boehm and Bonifacino, 2001; Robinson, 2004; Dhonukshe et al., 2007). Adaptor proteins (AP) play a key role (Boehm and Bonifacino, 2001; Robinson, 2004) in the recognition of sorting signals. APs are heterotetrameric protein complexes composed of two large subunits (β and γ/α/δ/ε), a small subunit (σ), and a medium subunit (µ) that is crucial for cargo selection (Boehm and Bonifacino, 2001). APs associate with the cytoplasmic side of secretory and endocytic vesicles, recruiting coat proteins and recognizing sorting signals within cargo proteins for their incorporation into vesicle carriers (Boehm and Bonifacino, 2001). Five APs have been identified so far, classified by their components, subcellular localization, and function (Boehm and Bonifacino, 2001; Robinson, 2004; Hirst et al., 2011). Of the five APs, AP-1 associates with the TGN or recycling endosomes (RE) in yeast and mammals (Huang et al., 2001; Robinson, 2004), mediating the sorting of cargo proteins to compartments of the endosomal-lysosomal system or to the basolateral PM of polarized epithelial cells (Gonzalez and Rodriguez-Boulan, 2009). Knockouts of AP-1 components in multicellular organisms resulted in embryonic lethality (Boehm and Bonifacino, 2001; Robinson, 2004).We show here that the recently identified Arabidopsis µ1 adaptin AP1M2 (Park et al., 2013; Teh et al., 2013) is a key component in the cellular machinery mediating intracellular protein sorting at the TGN/EE. AP1M2 was previously named HAPLESS13 (HAP13), whose mutant allele hap13 showed male gametophytic lethality (Johnson et al., 2004). In recent quests for AP-1 in plants, HAP13/AP1M2 was confirmed as the Arabidopsis µ1 adaptin based on its interaction with other components of the AP-1 complex as well as its localization at the TGN (Park et al., 2013; Teh et al., 2013). A novel mutant allele of HAP13/AP1M2, ap1m2-1, was found to be defective in the intracellular distribution of KNOLLE, leading to defective cytokinesis (Park et al., 2013; Teh et al., 2013). However, it was not clear whether HAP13/AP1M2 mediated other cellular activities and their developmental consequences. Using the same mutant allele, we found that functional loss of HAP13 (hap13-1/ap1m2-1) resulted in a full spectrum of growth defects, suggestive of compromised auxin signaling and of defective RLK signaling. Cell morphogenesis was also disturbed in hap13-1. Importantly, hap13-1 was insensitive to brefeldin A (BFA) washout, indicative of defects in guanine nucleotide exchange factors for ADP-ribosylation factor (ArfGEF)-mediated post-Golgi trafficking. Furthermore, HAP13/AP1M2 showed evolutionarily conserved function during vacuolar fusion, providing additional support to its identity as a µ1 adaptin. These results demonstrate the importance of the Arabidopsis µ1 adaptin for intracellular protein sorting centered on the TGN/EE.  相似文献   

13.
14.
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   

15.
16.
The halotolerant microalgae Dunaliella bardawil accumulates under nitrogen deprivation two types of lipid droplets: plastoglobuli rich in β-carotene (βC-plastoglobuli) and cytoplasmatic lipid droplets (CLDs). We describe the isolation, composition, and origin of these lipid droplets. Plastoglobuli contain β-carotene, phytoene, and galactolipids missing in CLDs. The two preparations contain different lipid-associated proteins: major lipid droplet protein in CLD and the Prorich carotene globule protein in βC-plastoglobuli. The compositions of triglyceride (TAG) molecular species, total fatty acids, and sn-1+3 and sn-2 positions in the two lipid pools are similar, except for a small increase in palmitic acid in plastoglobuli, suggesting a common origin. The formation of CLD TAG precedes that of βC-plastoglobuli, reaching a maximum after 48 h of nitrogen deprivation and then decreasing. Palmitic acid incorporation kinetics indicated that, at early stages of nitrogen deprivation, CLD TAG is synthesized mostly from newly formed fatty acids, whereas in βC-plastoglobuli, a large part of TAG is produced from fatty acids of preformed membrane lipids. Electron microscopic analyses revealed that CLDs adhere to chloroplast envelope membranes concomitant with appearance of small βC-plastoglobuli within the chloroplast. Based on these results, we propose that CLDs in D. bardawil are produced in the endoplasmatic reticulum, whereas βC-plastoglobuli are made, in part, from hydrolysis of chloroplast membrane lipids and in part, by a continual transfer of TAG or fatty acids derived from CLD.Eukaryotic cells accumulate neutral lipids in different tissues mainly in the form of lipid droplets (Murphy, 2012). Most lipid droplets consist of a core of triglycerides (TAGs) and/or sterol esters coated by a phospholipids monolayer and embedded with proteins (Zweytick et al., 2000). Plants accumulate TAGs in different tissues, primarily in seeds but also in fruit, such as palm oil, flowers, and leaves. The best characterized system for TAG metabolism is oil seeds, in which TAG serves as the major carbon and energy reservoir to be used during germination (Huang, 1992, 1996). Recent studies show that lipid droplets are not just static pools of lipids but have diverse metabolic functions (Farese and Walther, 2009). In addition, plants also contain plastoglobuli, small chloroplastic lipid droplets consisting primarily of storage lipids and pigments. Proteome analyses of plastoglobuli suggest that they are involved in synthesis and degradation of lipids, pigments, and coenzymes (Ytterberg et al., 2006; Lundquist et al., 2012). It has been shown that plant plastoglobuli are associated with thylakoid membranes (Austin et al., 2006; Ytterberg et al., 2006).It is not entirely clear where the TAGs are synthesized in the plant cell. Until recently, it has been assumed that most TAGs are made in the endoplasmatic reticulum (ER) from fatty acids, which are mostly synthesized in the chloroplast and imported to the cytoplasm (Joyard et al., 2010). However, the recent identification of the enzyme diacylglycerol acyl transferase in plant plastoglobuli (Lundquist et al., 2012) suggests that TAG may be synthesized directly in chloroplasts, although direct evidence is missing. TAG may be synthesized also from galactolipid fatty acids during stress or senescence by phytyl ester synthases, which catalyze acyl transesterification from galactolipids to TAGs (Lippold et al., 2012). Phosphatidyl choline (PC) plays a major role in acyl transfer of newly synthesized fatty acids from the chloroplast into TAGs at the ER in plants (Bates et al., 2009). An indication for the origin of glycerolipids in plants is the identity of the fatty acids at the sn-2 position: if it originates in the chloroplast, it is mostly C16:0, whereas if it was made in the ER, it is mostly C:18 (Heinz and Roughan, 1983).Many species of unicellular microalgae can accumulate large amounts of TAGs under growth-limiting conditions, such as nitrogen deprivation (Shifrin and Chisholm, 1981; Roessler, 1990; Avron and Ben-Amotz, 1992; Thompson, 1996). In green microalgae (Chlorophyceae), TAGs are usually synthesized and accumulated in cytoplasmatic lipid droplets (CLDs; Murphy, 2012), although in some cases, such as in Chlamydomonas reinhardtii starchless mutants, they also accumulate in chloroplasts (Fan et al., 2011; Goodson et al., 2011). Recent studies indicate that the CLDs are closely associated with ER membranes and possibly, chloroplast envelope membranes as well (Goodson et al., 2011; Peled et al., 2012).Green microalgae also contain two distinct types of chloroplastic lipid droplets. The first type is plastoglobuli, similar in morphology to higher plants plastoglobuli (Bréhélin et al., 2007; Kessler and Vidi, 2007). The second type is the eyespot (stigma), part of the visual system in microalgae. The eyespot is composed of a cluster of β-carotene-containing lipid droplets organized in several layers between grana membranes in the chloroplast (Häder and Lebert, 2009; Kreimer, 2009). Recent proteomic analysis of algal eyespot proteins revealed that they contain diverse structural proteins, lipid and carotenoid metabolizing enzymes, transporters, and signal transduction components (Schmidt et al., 2006).The origin of TAG in microalgae is still not clear. In C. reinhardtii, it was found that the major fatty acids in the sn-2 position are 16:0, which according to the plant dogma, is made in the chloroplast (Fan et al., 2011). In C. reinhardtii, which lacks PC, monogalactosyldiacylglycerol (MGDG) was proposed to replace PC in the mobilization of fatty acids from plastidal galactoglycerolipids into TAG based on mutation of a galactoglycerolipid lipase (Li et al., 2012). Based on these results and others, it has been proposed that, in C. reinhardtii, triglycerides are primarily produced in the chloroplast or combined with ER (Li et al., 2012; Liu and Benning, 2013).Plants and algae lipid droplets contain structural major proteins localized at the lipid droplet periphery, and their major function seems to be stabilization and prevention of fusion (Huang, 1992, 1996; Katz et al., 1995; Frandsen et al., 2001; Liu et al., 2009). In plant seed oils, the major classes of lipid droplet proteins are oleosins and caleosins, which have a characteristic hydrophobic loop with a conserved three Pro domain (Hsieh and Huang, 2004; Capuano et al., 2007; Purkrtova et al., 2008; Tzen, 2012). Oleosin and caleosin analogs were also recently identified in some green microalgal species (Lin et al., 2012; Vieler et al., 2012; Huang et al., 2013). However, the most abundant lipid droplets proteins in green algae (Chloropyceae) are a new family of major lipid droplet proteins (MLDPs) structurally distinct from plant oleosins and caleosins (Moellering and Benning, 2010; Peled et al., 2011; Davidi et al., 2012). Plastoglobules have different major lipid-associated proteins termed plastoglobules-associated protein-fibrillins, which form a distinct protein family with no sequence or structural similarities to oleosins (Kim and Huang, 2003). We have previously identified in the plastoglobuli rich in β-carotene (βC-plastoglobuli) a lipid-associated protein termed carotene globule protein (CGP), whose degradation destabilized the lipid droplets (Katz et al., 1995). The proteome of C. reinhardtii lipid droplet indicates that algal CLDs also contain several enzymes, suggesting that they are involved in lipid metabolism (Nguyen et al., 2011).The halotolerant green algae Dunaliella bardawil and Dunaliella salina ‘Teodoresco’ are unique in that they accumulate under high light stress or nitrogen deprivation large amounts of plastidic lipid droplets (βC-plastoglobuli), which consist of TAG and two isomers of β-carotene, all trans and 9-cis (Ben-Amotz et al., 1982, 1988). D. bardawil also accumulates CLD under the same stress conditions, similar to other green algae (Davidi et al., 2012). It has been shown that the function of βC-plastoglobuli is to protect the photosynthetic system against photoinhibition (Ben-Amotz et al., 1989). The enzymatic pathway for β-carotene synthesis in D. bardawil and D. salina has been partly identified, but the subcellular localization of β-carotene biosynthesis is not known (Jin and Polle, 2009). The synthesis of β-carotene depends on TAG biosynthesis (Rabbani et al., 1998); however, the origin of βC-plastoglobuli is not known. Are they formed within the chloroplast, or are they made in the cytoplasm? Is the TAG in βC-plastoglobuli and CLD identical or different, and where is it formed?D. bardawil is an excellent model organism for isolation of lipid droplet for several reasons. First, D. bardawil contains large amounts of both CLD and βC-plastoglobuli (Ben-Amotz et al., 1982; Fried et al., 1982), making it possible to obtain sufficient amounts of proteins and lipids from the two types of lipid pools for detailed analyses. Second, Dunaliella do not have a rigid cell wall and can be lysed by a gentle osmotic shock, which does not rupture the chloroplast. Therefore, it is possible to sequentially release pure CLD and βC-plastoglobuli by a two-step lysis (Katz et al., 1995). Third, D. bardawil seems to lack the eyespot structure, which can be clearly observed in other Dunaliella spp. even in a light microscope or by electron microscopy, but has never been observed in D. bardawil by us. It avoids the risk of cross contamination of βC-plastoglobuli with eyespot proteins. Fourth, the availability of protein markers for the major lipid droplet-associated proteins, CGPs and MLDPs, enabled both good immunolocalization and careful monitoring of the purity of the preparations by western analysis.In this work, we describe the purification, lipid compositions, and protein profiles of two lipid pools from D. bardawil: CLD and plastidic βC-plastoglobuli. A detailed proteomic analysis of these lipid droplets will be described in another work. Combined with detailed electron microscopy studies, these results led to surprising conclusions regarding the origin of the plastidic βC-plastoglobuli.  相似文献   

17.
In rice (Oryza sativa) roots, lysigenous aerenchyma, which is created by programmed cell death and lysis of cortical cells, is constitutively formed under aerobic conditions, and its formation is further induced under oxygen-deficient conditions. Ethylene is involved in the induction of aerenchyma formation. reduced culm number1 (rcn1) is a rice mutant in which the gene encoding the ATP-binding cassette transporter RCN1/OsABCG5 is defective. Here, we report that the induction of aerenchyma formation was reduced in roots of rcn1 grown in stagnant deoxygenated nutrient solution (i.e. under stagnant conditions, which mimic oxygen-deficient conditions in waterlogged soils). 1-Aminocyclopropane-1-carboxylic acid synthase (ACS) is a key enzyme in ethylene biosynthesis. Stagnant conditions hardly induced the expression of ACS1 in rcn1 roots, resulting in low ethylene production in the roots. Accumulation of saturated very-long-chain fatty acids (VLCFAs) of 24, 26, and 28 carbons was reduced in rcn1 roots. Exogenously supplied VLCFA (26 carbons) increased the expression level of ACS1 and induced aerenchyma formation in rcn1 roots. Moreover, in rice lines in which the gene encoding a fatty acid elongase, CUT1-LIKE (CUT1L; a homolog of the gene encoding Arabidopsis CUT1, which is required for cuticular wax production), was silenced, both ACS1 expression and aerenchyma formation were reduced. Interestingly, the expression of ACS1, CUT1L, and RCN1/OsABCG5 was induced predominantly in the outer part of roots under stagnant conditions. These results suggest that, in rice under oxygen-deficient conditions, VLCFAs increase ethylene production by promoting 1-aminocyclopropane-1-carboxylic acid biosynthesis in the outer part of roots, which, in turn, induces aerenchyma formation in the root cortex.Aerenchyma formation is a morphological adaptation of plants to complete submergence and waterlogging of the soil, and facilitates internal gas diffusion (Armstrong, 1979; Jackson and Armstrong, 1999; Colmer, 2003; Voesenek et al., 2006; Bailey-Serres and Voesenek, 2008; Licausi and Perata, 2009; Sauter, 2013; Voesenek and Bailey-Serres, 2015). To adapt to waterlogging in soil, rice (Oryza sativa) develops lysigenous aerenchyma in shoots (Matsukura et al., 2000; Colmer and Pedersen, 2008; Steffens et al., 2011) and roots (Jackson et al., 1985b; Justin and Armstrong, 1991; Kawai et al., 1998), which is formed by programmed cell death and subsequent lysis of some cortical cells (Jackson and Armstrong, 1999; Evans, 2004; Yamauchi et al., 2013). In rice roots, lysigenous aerenchyma is constitutively formed under aerobic conditions (Jackson et al., 1985b), and its formation is further induced under oxygen-deficient conditions (Colmer et al., 2006; Shiono et al., 2011). The former and latter are designated constitutive and inducible lysigenous aerenchyma formation, respectively (Colmer and Voesenek, 2009). The gaseous plant hormone ethylene regulates adaptive growth responses of plants to submergence (Voesenek and Blom, 1989; Voesenek et al., 1993; Visser et al., 1996a,b; Lorbiecke and Sauter, 1999; Hattori et al., 2009; Steffens and Sauter, 2009; van Veen et al., 2013). Ethylene also induces lysigenous aerenchyma formation in roots of some gramineous plants (Drew et al., 2000; Shiono et al., 2008). The treatment of roots with ethylene or its precursor (1-aminocyclopropane-1-carboxylic acid [ACC]) stimulates aerenchyma formation in rice (Justin and Armstrong, 1991; Colmer et al., 2006; Yukiyoshi and Karahara, 2014), maize (Zea mays; Drew et al., 1981; Jackson et al., 1985a; Takahashi et al., 2015), and wheat (Triticum aestivum; Yamauchi et al., 2014a,b). Moreover, treatment of roots with inhibitors of ethylene action or ethylene biosynthesis effectively blocks aerenchyma formation under hypoxic conditions in maize (Drew et al., 1981; Konings, 1982; Jackson et al., 1985a; Rajhi et al., 2011).Ethylene biosynthesis is accomplished by two main successive enzymatic reactions: conversion of S-adenosyl-Met to ACC by 1-aminocyclopropane-1-carboxylic acid synthase (ACS), and conversion of ACC to ethylene by 1-aminocyclopropane-1-carboxylic acid oxidase (ACO; Yang and Hoffman, 1984). The activities of both enzymes are enhanced during aerenchyma formation under hypoxic conditions in maize root (He et al., 1996). Since the ACC content in roots of maize is increased by oxygen deficiency and is strongly correlated with ethylene production (Atwell et al., 1988), ACC biosynthesis is essential for ethylene production during aerenchyma formation in roots. In fact, exogenously supplied ACC induced ethylene production in roots of maize (Drew et al., 1979; Konings, 1982; Atwell et al., 1988) and wheat (Yamauchi et al., 2014b), even under aerobic conditions. Ethylene production in plants is inversely related to oxygen concentration (Yang and Hoffman, 1984). Under anoxic conditions, the oxidation of ACC to ethylene by ACO, which requires oxygen, is almost completely repressed (Yip et al., 1988; Tonutti and Ramina, 1991). Indeed, anoxic conditions stimulate neither ethylene production nor aerenchyma formation in maize adventitious roots (Drew et al., 1979). Therefore, it is unlikely that the root tissues forming inducible aerenchyma are anoxic, and that the ACO-mediated step is repressed. Moreover, aerenchyma is constitutively formed in rice roots even under aerobic conditions (Jackson et al., 1985b), and thus, after the onset of waterlogging, oxygen can be immediately supplied to the apical regions of roots through the constitutively formed aerenchyma.Very-long-chain fatty acids (VLCFAs; ≥20 carbons) are major constituents of sphingolipids, cuticular waxes, and suberin in plants (Franke and Schreiber, 2007; Kunst and Samuels, 2009). In addition to their structural functions, VLCFAs directly or indirectly participate in several physiological processes (Zheng et al., 2005; Reina-Pinto et al., 2009; Roudier et al., 2010; Ito et al., 2011; Nobusawa et al., 2013; Tsuda et al., 2013), including the regulation of ethylene biosynthesis (Qin et al., 2007). During fiber cell elongation in cotton ovules, ethylene biosynthesis is enhanced by treatment with saturated VLCFAs, especially 24-carbon fatty acids, and is suppressed by an inhibitor of VLCFA biosynthesis (Qin et al., 2007). The first rate-limiting step in VLCFA biosynthesis is condensation of acyl-CoA with malonyl-CoA by β-ketoacyl-CoA synthase (KCS; Joubès et al., 2008). KCS enzymes are thought to determine the substrate and tissue specificities of fatty acid elongation (Joubès et al., 2008). The Arabidopsis (Arabidopsis thaliana) genome has 21 KCS genes (Joubès et al., 2008). In the Arabidopsis cut1 mutant, which has a defect in the gene encoding CUT1 that is required for cuticular wax production (i.e. one of the KCS genes), the expression of AtACO genes and growth of root cells were reduced when compared with the wild type (Qin et al., 2007). Furthermore, expression of the AtACO genes was rescued by exogenously supplied saturated VLCFAs (Qin et al., 2007). These observations imply that VLCFAs or their derivatives work as regulatory factors for gene expression during some physiological processes in plants.reduced culm number1 (rcn1) was first identified as a rice mutant with a low tillering rate in a paddy field (Takamure and Kinoshita, 1985; Yasuno et al., 2007). The rcn1 (rcn1-2) mutant has a single nucleotide substitution in the gene encoding a member of the ATP-binding cassette (ABC) transporter subfamily G, RCN1/OsABCG5, causing an Ala-684Pro substitution (Yasuno et al., 2009). The mutation results in several mutant phenotypes, although the substrates of RCN1/OsABCG5 have not been determined (Ureshi et al., 2012; Funabiki et al., 2013; Matsuda et al., 2014). We previously found that the rcn1 mutant has abnormal root morphology, such as shorter root length and brownish appearance of roots, under stagnant (deoxygenated) conditions (which mimics oxygen-deficient conditions in waterlogged soils). We also found that the rcn1 mutant accumulates less of the major suberin monomers originating from VLCFAs in the outer part of adventitious roots, and this results in a reduction of a functional apoplastic barrier in the root hypodermis (Shiono et al., 2014a).The objective of this study was to elucidate the molecular basis of inducible aerenchyma formation. To this end, we examined lysigenous aerenchyma formation and ACC, ethylene, and VLCFA accumulation and their biosyntheses in rcn1 roots. Based on the results of these studies, we propose that VLCFAs are involved in inducible aerenchyma formation through the enhancement of ethylene biosynthesis in rice roots.  相似文献   

18.
19.
Proteins decorated with arabinogalactan (AG) have important roles in cell wall structure and plant development, yet the structure and biosynthesis of this polysaccharide are poorly understood. To facilitate the analysis of biosynthetic mutants, water-extractable arabinogalactan proteins (AGPs) were isolated from the leaves of Arabidopsis (Arabidopsis thaliana) plants and the structure of the AG carbohydrate component was studied. Enzymes able to hydrolyze specifically AG were utilized to release AG oligosaccharides. The released oligosaccharides were characterized by high-energy matrix-assisted laser desorption ionization-collision-induced dissociation mass spectrometry and polysaccharide analysis by carbohydrate gel electrophoresis. The Arabidopsis AG is composed of a β-(1→3)-galactan backbone with β-(1→6)-d-galactan side chains. The β-(1→6)-galactan side chains vary in length from one to over 20 galactosyl residues, and they are partly substituted with single α-(1→3)-l-arabinofuranosyl residues. Additionally, a substantial proportion of the β-(1→6)-galactan side chain oligosaccharides are substituted at the nonreducing termini with single 4-O-methyl-glucuronosyl residues via β-(1→6)-linkages. The β-(1→6)-galactan side chains are occasionally substituted with α-l-fucosyl. In the fucose-deficient murus1 mutant, AGPs lack these fucose modifications. This work demonstrates that Arabidopsis mutants in AGP structure can be identified and characterized. The detailed structural elucidation of the AG polysaccharides from the leaves of Arabidopsis is essential for insights into the structure-function relationships of these molecules and will assist studies on their biosynthesis.Arabinogalactans (AGs) are structurally complex large-branched polysaccharides attached to Hyp residues of many plant cell wall polypeptides. Most proteins glycosylated with AGs (AGPs) have both AG glycosylated domains (glycomodules) and structural or enzymatic domains. However, typical AGPs commonly contain less than 10% protein, suggesting that the AG is the functional part of the molecule (Clarke et al., 1979; Fincher et al., 1983; Kieliszewski and Lamport, 1994; Borner et al., 2003; Xu et al., 2008). Hyp is the most characteristic amino acid present at the glycosylated domain of the AGP, but other amino acids such as Ser, Ala, and Thr are also very common. Type II AG polysaccharides share common structural features based on a β-(1→3)-galactan backbone with β-(1→6)-linked galactan side chains and can be found both on AGPs and rhamnogalacturonan-I (RG-I) pectin (Renard et al., 1991). The galactopyranosyl (Galp) residues can be further substituted with l-arabinofuranosyl (l-Araf) and occasionally also l-rhamnosyl (l-Rha), l-fucosyl (l-Fuc), and glucuronosyl (GlcA; with or without 4-O-methylation) residues (Tsumuraya et al., 1988; Tan et al., 2004; Tryfona et al., 2010). (Sugars mentioned in this work belong to the D-series unless otherwise stated.)The structure of AGs is poorly characterized, and this is mainly due to the great heterogeneity of glycan structures, not only between different AGPs but also even on the same peptide sequence in the same tissue (Estévez et al., 2006). The glycan structure can also be different depending on the developmental stage and tissue type (Tsumuraya et al., 1988), adding to the great heterogeneity of these molecules and therefore limiting their detailed characterization. Molecular and biochemical evidence has indicated that AGPs have specific functions during root formation, promotion of somatic embryogenesis (van Hengel et al., 2002), and attraction of pollen tubes to the style (Cheung et al., 1995). In addition, enhanced secretion efficiency or stability in the cell wall are properties that the AG may confer on the glycosylated protein (Borner et al., 2003). However, it has been difficult to differentiate one species of AGP from another in plant tissues and to assign specific roles to individual AGPs.l-Fuc is present in AGPs in Arabidopsis (Arabidopsis thaliana; van Hengel et al., 2002), radish (Raphanus sativus; Nakamura et al., 1984; Tsumuraya et al., 1984a, 1984b, 1988), and several other dicot plants such as thyme (Thymus vulgaris; Chun et al., 2001) and celery (Apium graveolens; Lin et al., 2011). Reduction in l-Fuc by 40% in roots of murus1 (mur1) plants resulted in a decrease of 50% in root cell elongation, and eel lectin binding assays suggested that the phenotype was the result of alterations in the composition of root AGPs (van Hengel and Roberts, 2002). An α-(1→2)-fucosyltransferase (FUT) activity for radish primary root AGPs has been described, where an α-l-Araf-(1→3)-β-Galp-(1→6)-Galp trisaccharide was used as exogenous substrate acceptor to mimic an AG polysaccharide in the enzymatic assay (Misawa et al., 1996). Linkage analysis, reactivity with eel lectin, and digestion with α-(1→2)-fucosidase indicated that the l-Fuc residues added are terminal and attached via an α-linkage to the C-2 position of an adjacent l-Araf residue (Nakamura et al., 1984; Tsumuraya et al., 1984a, 1984b, 1988). Recently, Wu et al. (2010) identified AtFUT4 and AtFUT6 genes encoding FUT proteins specific to AGPs, but the structures of the fucosylated AG generated have not been fully characterized.To gain insights into the synthesis and function of plant AGPs, it would be useful to have mutants altered in their carbohydrate moieties. However, no AG-specific biosynthetic mutants have been characterized, and this, among other reasons, is due to the very limited knowledge of the structure of Arabidopsis AGs (Qu et al., 2008). Moreover, characterization of AG in candidate mutants remains challenging. Even though the structures of some AGs have been proposed using NMR and sugar linkage analyses, the complete structural elucidation of a native AG still remains a formidable task, because NMR spectroscopy and methylation analysis have been largely used to provide information regarding the amount and type of linkages between adjacent glycosyl residues, and AG heterogeneity can confound attempts to build complete structural models. Recently, a modular structure was proposed for AGs on heterologously expressed proteins in tobacco (Nicotiana tabacum; Tan et al., 2010). Tan et al. (2010) proposed that approximately 15-residue repeating blocks of decorated β-(1→3)-trigalactosyl subunits connected by β-(1→6)-linkages were the building blocks of type II AG polysaccharides and concluded that these molecules are far less complex than commonly supposed. Most characterized β-(1→6)-galactan side chains in AGs are reported to be short, of one or two residues (Neukom and Markwalder, 1975; Gane et al., 1995; Gaspar et al., 2001). On the contrary, there are reports of long β-(1→6)-galactan side chains in radish root AGPs (Haque et al., 2005). Similarly, we recently found evidence that wheat (Triticum aestivum) flour endosperm AGP extracts contained long β-(1→6)-galactan side chains heavily substituted with l-Araf at C-3 (Tryfona et al., 2010). This partial structure of the carbohydrate component of wheat flour AGP isolated from water extracts of wheat endosperm was elucidated utilizing a combination of analytical approaches, such as the use of enzymes able to release oligosaccharides specifically from AGs, high-energy matrix-assisted laser desorption ionization (MALDI)-collision-induced dissociation (CID) mass spectrometry (MS), and polysaccharide analysis by carbohydrate gel electrophoresis (PACE; Tryfona et al., 2010). In this work, we applied these techniques to study the carbohydrate component of Arabidopsis leaf AGPs. AG-specific enzyme digestion products were analyzed by PACE and MS, allowing a partial structure to be proposed. We show that endogenous Arabidopsis leaf AG is composed of a β-(1→3)-galactan backbone with β-(1→6)-galactan side chains. These side chains are substituted with l-Araf residues via α-(1→3)-linkages and can vary in length from one up to at least 20 Galp residues. We also found that the β-(1→6)-galactan side chains are substituted mainly with 4-O-methyl-glucuronosyl (4-O-Me-GlcA) at their nonreducing termini, while occasional l-Fuc substitutions were also present via α-(1→2)-linkages on l-Araf residues. In addition, AG oligosaccharides from leaves of the mur1 mutant were identified, and their structures were compared with those isolated from wild-type plants.  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