首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
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   

3.
4.
5.
Identification of mechanisms that decrease cadmium accumulation in plants is a prerequisite for minimizing dietary uptake of cadmium from contaminated crops. Here, we show that cadmium inhibits nitrate transporter 1.1 (NRT1.1)-mediated nitrate (NO3) uptake in Arabidopsis (Arabidopsis thaliana) and impairs NO3 homeostasis in roots. In NO3-containing medium, loss of NRT1.1 function in nrt1.1 mutants leads to decreased levels of cadmium and several other metals in both roots and shoots and results in better biomass production in the presence of cadmium, whereas in NO3-free medium, no difference is seen between nrt1.1 mutants and wild-type plants. These results suggest that inhibition of NRT1.1 activity reduces cadmium uptake, thus enhancing cadmium tolerance in an NO3 uptake-dependent manner. Furthermore, using a treatment rotation system allowing synchronous uptake of NO3 and nutrient cations and asynchronous uptake of cadmium, the nrt1.1 mutants had similar cadmium levels to wild-type plants but lower levels of nutrient metals, whereas the opposite effect was seen using treatment rotation allowing synchronous uptake of NO3 and cadmium and asynchronous uptake of nutrient cations. We conclude that, although inhibition of NRT1.1-mediated NO3 uptake by cadmium might have negative effects on nitrogen nutrition in plants, it has a positive effect on cadmium detoxification by reducing cadmium entry into roots. NRT1.1 may regulate the uptake of cadmium and other cations by a common mechanism.Cadmium is highly toxic to humans (Nicholson et al., 1983), and its primary route of entry into the body is through crops grown in cadmium-contaminated soil (Clemens et al., 2013). A recent survey indicated that vegetables and rice (Oryza sativa) account for approximately 40% and 38%, respectively, of total cadmium exposure in residents of Shanghai, China’s largest city (He et al., 2013). However, cadmium contamination of agricultural soils as a result of rapid industrial development and release of agrochemicals into the environment is an increasingly serious problem. Many strategies have been proposed for remediating cadmium-contaminated soil to prevent cadmium uptake by crops. These strategies include the dig-and-dump method or encapsulation of the contaminated soil, chemical immobilization or extraction of cadmium, and phytoremediation by cadmium-hyperaccumulating plants (Pulford and Watson, 2003). However, the dig-and-dump and chemical methods are expensive, whereas phytoremediation requires several growing seasons to be effective, making it impractical in regions where farmland is limited and food supply insufficient.The shortfalls of these strategies have prompted researchers to develop alternative techniques that are cost-effective and interfere less with crop production. Use of nitrogen fertilizers is one of the most important agronomic practices and it has been suggested that their appropriate use might provide a relatively inexpensive, time-saving, and effective strategy for reducing cadmium entry into, and accumulation in, crops because NO3 facilitates cadmium uptake in hydroponically grown plants (Sarwar et al., 2010; Luo et al., 2012). However, in a preliminary study, we found that, in plants grown in soil, the effect of the nitrogen form on cadmium accumulation was strongly associated with the pH-buffering capacity of the soil. In soil with a lower pH-buffering capacity, application of ammonium (NH4+) resulted in higher cadmium levels in plants than application of NO3, probably as a result of soil acidification by NH4+, and the opposite effect was seen when plants were grown in soil with higher pH-buffering capacity (S.K. Fan, S.T. Du, and C.W. Jin, unpublished data). This suggests that management of the use of nitrogen fertilizers to prevent cadmium entry into crops might be difficult because of the wide variation in soil pH-buffering capacity.Because NO3 facilitates cadmium uptake in hydroponically grown plants as described above, modification of NO3 uptake pathways in plants might also affect cadmium uptake, in which case modifying these pathways to reduce cadmium entry into crops could circumvent the risks and the difficulties involved in nitrogen fertilizer management. Exposure to cadmium has been shown to reduce NO3 uptake by roots (Hernández et al., 1997; Gouia et al., 2000; Rizzardo et al., 2012), but this has been assumed to be deleterious to plant growth (Finkemeier et al., 2003; Rizzardo et al., 2012). The process by which NO3 is taken up across the root plasma membrane is complex, and several nitrate transporters (NRTs) involved in NO3 uptake from the growth medium have been characterized. In Arabidopsis (Arabidopsis thaliana), NRT1.1 is a dual-affinity transporter involved in both high- and low-affinity uptake, NRT1.2 is involved only in low-affinity NO3 uptake, whereas NRT2.1, NRT2.2, and NRT2.4 are only involved in high-affinity NO3 uptake (Wang et al., 2012; Léran et al., 2014). However, the transporter responsible for the cadmium-induced decrease in NO3 uptake remains unknown. Given the presumed association between NO3 uptake and cadmium uptake, it is important to identify the molecular mechanism involved in this process, and it is particularly important to determine whether the modulation of relevant NO3 transporters affects cadmium entry into plants.In this study, we investigated the relationship between NO3 uptake and cadmium uptake in Arabidopsis roots. To our knowledge, our results reveal a new mechanism for resisting cadmium toxicity: Cadmium reduces NO3 uptake by inhibiting NRT1.1 activity, which in turn reduces cadmium entry into root cells. As a result, cadmium levels in plants are lower and plant growth is improved. Our findings may provide a strategy for minimizing cadmium accumulation in crops grown in contaminated soil using biotechnological pathways to decrease NO3 uptake.  相似文献   

6.
The Kv-like (potassium voltage-dependent) K+ channels at the plasma membrane, including the inward-rectifying KAT1 K+ channel of Arabidopsis (Arabidopsis thaliana), are important targets for manipulating K+ homeostasis in plants. Gating modification, especially, has been identified as a promising means by which to engineer plants with improved characteristics in mineral and water use. Understanding plant K+ channel gating poses several challenges, despite many similarities to that of mammalian Kv and Shaker channel models. We have used site-directed mutagenesis to explore residues that are thought to form two electrostatic countercharge centers on either side of a conserved phenylalanine (Phe) residue within the S2 and S3 α-helices of the voltage sensor domain (VSD) of Kv channels. Consistent with molecular dynamic simulations of KAT1, we show that the voltage dependence of the channel gate is highly sensitive to manipulations affecting these residues. Mutations of the central Phe residue favored the closed KAT1 channel, whereas mutations affecting the countercharge centers favored the open channel. Modeling of the macroscopic current kinetics also highlighted a substantial difference between the two sets of mutations. We interpret these findings in the context of the effects on hydration of amino acid residues within the VSD and with an inherent bias of the VSD, when hydrated around a central Phe residue, to the closed state of the channel.Plant cells utilize the potassium ion (K+) to maintain hydrostatic (turgor) pressure, to drive irreversible cell expansion for growth, and to facilitate reversible changes in cell volume during stomatal movements. Potassium uptake and its circulation throughout the plant relies both on high-affinity, H+-coupled K+ transport (Quintero and Blatt, 1997; Rubio et al., 2008) and on K+ channels to facilitate K+ ion transfer across cell membranes. Uptake via K+ channels is thought to be responsible for roughly 50% of the total K+ content of the plant under most field conditions (Spalding et al., 1999; Rubio et al., 2008; Amtmann and Blatt, 2009). K+ channels confer on the membranes of virtually every tissue distinct K+ conductances and regulatory characteristics (Véry and Sentenac, 2003; Dreyer and Blatt, 2009). Their characteristics are thus of interest for engineering directed to manipulating K+ flux in many aspects of plant growth and cellular homeostasis. The control of K+ channel gating has been identified as the most promising target for the genetic engineering of stomatal responsiveness (Lawson and Blatt, 2014; Wang et al., 2014a), based on the recent development of quantitative systems models of guard cell transport and metabolism (Chen et al., 2012b; Hills et al., 2012; Wang et al., 2012). By contrast, modifying the expression and, most likely, the population of native K+ channels at the membrane was found to have no substantial effect on stomatal physiology (Wang et al., 2014b).The Kv-like K+ channels of the plant plasma membrane (Pilot et al., 2003; Dreyer and Blatt, 2009) share a number of structural features with the Kv superfamily of K+ channels characterized in animals and Drosophila melanogaster (Papazian et al., 1987; Pongs et al., 1988). The functional channels assemble from four homologous subunits and surround a central transmembrane pore that forms the permeation pathway (Daram et al., 1997). Each subunit comprises six transmembrane α-helices, designated S1 to S6, and both N and C termini are situated on the cytosolic side of the membrane (Uozumi et al., 1998). The pore or P loop between the S5 and S6 α-helices incorporates a short α-helical stretch and the highly conserved amino acid sequence TxGYGD, which forms a selectivity filter for K+ (Uozumi et al., 1995; Becker et al., 1996; Nakamura et al., 1997). The carbonyl oxygen atoms of these residues in all four K+ channel subunits face inward to form coordination sites for K+ ions between them (Doyle et al., 1998; Jiang et al., 2003; Kuo et al., 2003; Long et al., 2005) and a multiple-ion pore (Thiel and Blatt, 1991) such that K+ ions pass through the selectivity filter as if in free solution. The plant channels are also sensitive to a class of neurotoxins that exhibit high specificity in binding around the mouth of the channel pore (Obermeyer et al., 1994).These K+ channels also share a common gating mechanism. Within each subunit, the first four α-helices form a quasiindependent unit, the voltage sensor domain (VSD), with the S4 α-helix incorporating positively charged (Arg or Lys) residues regularly positioned across the lipid bilayer and transmembrane electric field. Voltage displaces the S4 α-helix within the membrane and couples rotation of the S5 and S6 α-helices lining the pore, thereby opening or closing the channel (Sigworth, 2003; Dreyer and Blatt, 2009). For outward-rectifying channels, such as the mammalian Kv1.2 and the D. melanogaster Shaker K+ channels, an inside-positive electric field drives the positively charged, S4 α-helix outward (the up position), which draws on the S4-S5 linker to open the pore. This simple expedient of a lever and string secures current flow in one direction by favoring opening at positive, but not negative, voltages. This same model applies to the Arabidopsis (Arabidopsis thaliana) Kv-like K+ channels, including outward rectifiers that exhibit sensitivity to external K+ concentration (Blatt, 1988; Blatt and Gradmann, 1997; Johansson et al., 2006), and it serves equally in the gating of inward-rectifying K+ channels such as KAT1, which gates open at negative voltages (Dreyer and Blatt, 2009).Studies of KAT1 gating (Latorre et al., 2003; Lai et al., 2005) have indicated that the S4 α-helix of the channel most likely undergoes very similar conformational changes with voltage as those of the mammalian and Shaker K+ channels. These findings conform with the present understanding of the evolution of VSD structure (Palovcak et al., 2014) and the view of a common functional dynamic to its molecular design. It is likely, therefore, that a similar electrostatic network occurs in KAT1 to stabilize the VSD. Crucially, however, experimental evidence in support of such a network has yet to surface. Electrostatic countercharges and the hydration of amino acid side chains between the α-helices within the VSDs of mammalian and Shaker K+ channel models are important for the latch-like stabilization of the so-called down and up states of these channels (Tao et al., 2010; Pless et al., 2011). Nonetheless, some studies (Gajdanowicz et al., 2009; Riedelsberger et al., 2010) have pointed to subtle differences in the structure of KAT1 that relate to the VSD.We have explored the electrostatic network of the KAT1 VSD through site-directed mutagenesis to manipulate the voltage dependence of KAT1, combining these studies with molecular dynamic simulations previously shown to accommodate the plant VSDs and their hydration during gating transitions (Gajdanowicz et al., 2009; Garcia-Mata et al., 2010). We report here that gating of KAT1 is sensitive to manipulations affecting a set of electrostatic charge transfer centers. These findings conform in large measure to the mammalian and Shaker models. However, virtually all manipulations affecting a highly conserved, central Phe favor the up state of the VSD and the closed KAT1 channel, whereas mutations affecting the electrostatic networks on either side of this Phe favor the down state of the VSD and the open channel. These and additional observations suggest that hydration within the VSD is a major determinant of KAT1 gating.  相似文献   

7.
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   

8.
9.
10.
Hydrogen sulfide (H2S) is the third biological gasotransmitter, and in animals, it affects many physiological processes by modulating ion channels. H2S has been reported to protect plants from oxidative stress in diverse physiological responses. H2S closes stomata, but the underlying mechanism remains elusive. Here, we report the selective inactivation of current carried by inward-rectifying K+ channels of tobacco (Nicotiana tabacum) guard cells and show its close parallel with stomatal closure evoked by submicromolar concentrations of H2S. Experiments to scavenge H2S suggested an effect that is separable from that of abscisic acid, which is associated with water stress. Thus, H2S seems to define a unique and unresolved signaling pathway that selectively targets inward-rectifying K+ channels.Hydrogen sulfide (H2S) is a small bioactive gas that has been known for centuries as an environmental pollutant (Reiffenstein et al., 1992). H2S is soluble in both polar and, especially, nonpolar solvents (Wang, 2002), and has recently come to be recognized as the third member of a group of so-called biological gasotransmitters. Most importantly, H2S shows both physical and functional similarities to the other gasotransmitters nitric oxide (NO) and carbon monoxide (Wang, 2002), and it has been shown to participate in diverse physiological processes in animals, including cardioprotection, neuromodulation, inflammation, apoptosis, and gastrointestinal functions among others (Kabil et al., 2014). Less is known about H2S molecular targets and its modes of action. H2S can directly modify specific targets through protein sulfhydration (the addition of an -SH group to thiol moiety of proteins; Mustafa et al., 2009) or reaction with metal centers (Li and Lancaster, 2013). It can also act indirectly, reacting with NO to form nitrosothiols (Whiteman et al., 2006; Li and Lancaster, 2013). Among its molecular targets, H2S has been reported to regulate ATP-dependent K+ channels (Yang et al., 2005), Ca2+-activated K+ channels, T- and L-type Ca2+ channels, and transient receptor potential channels (Tang et al., 2010; Peers et al., 2012), suggesting H2S as a key regulator of membrane ion transport.In plants, H2S is produced enzymatically by the desulfhydration of l-Cys to form H2S, pyruvate, and ammonia in a reaction catalyzed by the enzyme l-Cys desulfhydrase (Riemenschneider et al., 2005a, 2005b), DES1, that has been characterized in Arabidopsis (Arabidopsis thaliana; Alvarez et al., 2010). Alternatively, H2S can be produced from d-Cys by d-Cys desulfhydrase (Riemenschneider et al., 2005a, 2005b) and in cyanide metabolism by β-cyano-Ala synthase (García et al., 2010). H2S action was originally related to pathogenesis resistance (Bloem et al., 2004), but in the last decade it has been proven to have an active role in signaling, participating in key physiological processes, such as germination and root organogenesis (Zhang et al., 2008, 2009a), heat stress (Li et al., 2013a, 2013b), osmotic stress (Zhang et al., 2009b), and stomatal movement (García-Mata and Lamattina, 2010; Lisjak et al., 2010, 2011; Jin et al., 2013). Moreover, H2S was reported to participate in the signaling of plant hormones, including abscisic acid (ABA; García-Mata and Lamattina, 2010; Lisjak et al., 2010; Jin et al., 2013; Scuffi et al., 2014), ethylene (Hou et al., 2013), and auxin (Zhang et al., 2009a).ABA is an important player in plant physiology. Notably, upon water stress, ABA triggers a complex signaling network to restrict the loss of water through the transpiration stream, balancing these needs with those of CO2 for carbon assimilation. In the guard cells that surround the stomatal pore, ABA induces an increase of cytosolic-free Ca2+ concentration ([Ca2+]cyt), elevates cytosolic pH (pHi), and activates the efflux of anions, mainly chloride, through S- and R-type anion channels. The increase in [Ca2+]cyt inactivates inward-rectifying K+ channels (IKIN); anion efflux depolarizes the plasma membrane, and together with the rise in pHi, it activates K+ efflux through outward-rectifying K+ channels (IKOUT; Blatt, 2000; Schroeder et al., 2001). These changes in ion flux, in turn, generate an osmotically driven reduction in turgor and volume and closure of the stomatal pore. All three gasotransmitters have been implicated in regulating the activity of guard cell ion channels, but direct evidence is available only for NO (Garcia-Mata et al., 2003; Sokolovski et al., 2005). Here, we have used two-electrode voltage clamp measurements to study the role of H2S in the regulation of the guard cell K+ channels of tobacco (Nicotiana tabacum). Our results show that H2S selectively inactivates IKIN and that this action parallels that of stomatal closure. These results confirm H2S as a unique factor regulating guard cell ion transport and indicate that H2S acts in a manner separable from that of ABA.  相似文献   

11.
In plants, K transporter (KT)/high affinity K transporter (HAK)/K uptake permease (KUP) is the largest potassium (K) transporter family; however, few of the members have had their physiological functions characterized in planta. Here, we studied OsHAK5 of the KT/HAK/KUP family in rice (Oryza sativa). We determined its cellular and tissue localization and analyzed its functions in rice using both OsHAK5 knockout mutants and overexpression lines in three genetic backgrounds. A β-glucuronidase reporter driven by the OsHAK5 native promoter indicated OsHAK5 expression in various tissue organs from root to seed, abundantly in root epidermis and stele, the vascular tissues, and mesophyll cells. Net K influx rate in roots and K transport from roots to aerial parts were severely impaired by OsHAK5 knockout but increased by OsHAK5 overexpression in 0.1 and 0.3 mm K external solution. The contribution of OsHAK5 to K mobilization within the rice plant was confirmed further by the change of K concentration in the xylem sap and K distribution in the transgenic lines when K was removed completely from the external solution. Overexpression of OsHAK5 increased the K-sodium concentration ratio in the shoots and salt stress tolerance (shoot growth), while knockout of OsHAK5 decreased the K-sodium concentration ratio in the shoots, resulting in sensitivity to salt stress. Taken together, these results demonstrate that OsHAK5 plays a major role in K acquisition by roots faced with low external K and in K upward transport from roots to shoots in K-deficient rice plants.Potassium (K) is one of the three most important macronutrients and the most abundant cation in plants. As a major osmoticum in the vacuole, K drives the generation of turgor pressure, enabling cell expansion. In the vascular tissue, K is an important participant in the generation of root pressure (for review, see Wegner, 2014 [including his new hypothesis]). In the phloem, K is critical for the transport of photoassimilates from source to sink (Marschner, 1996; Deeken et al., 2002; Gajdanowicz et al., 2011). In addition, enhancing K absorption and decreasing sodium (Na) accumulation is a major strategy of glycophytes in salt stress tolerance (Maathuis and Amtmann, 1999; Munns and Tester, 2008; Shabala and Cuin, 2008).Plants acquire K through K-permeable proteins at the root surface. Since available K concentration in the soil may vary by 100-fold, plants have developed multiple K uptake systems for adapting to this variability (Epstein et al., 1963; Grabov, 2007; Maathuis, 2009). In a classic K uptake experiment in barley (Hordeum vulgare), root K absorption has been described as a high-affinity and low-affinity biphasic transport process (Epstein et al., 1963). It is generally assumed that the low-affinity transport system (LATS) in the roots mediates K uptake in the millimolar range and that the activity of this system is insensitive to external K concentration (Maathuis and Sanders, 1997; Chérel et al., 2014). In contrast, the high-affinity transport system (HATS) was rapidly up-regulated when the supply of exogenous K was halted (Glass, 1976; Glass and Dunlop, 1978).The membrane transporters for K flux identified in plants are generally classified into three channels and three transporter families based on phylogenetic analysis (Mäser et al., 2001; Véry and Sentenac, 2003; Lebaudy et al., 2007; Alemán et al., 2011). For K uptake, it was predicted that, under most circumstances, K transporters function as HATS, while K-permeable channels mediate LATS (Maathuis and Sanders, 1997). However, a root-expressed K channel in Arabidopsis (Arabidopsis thaliana), Arabidopsis K Transporter1 (AKT1), mediates K absorption over a wide range of external K concentrations (Sentenac et al., 1992; Lagarde et al., 1996; Hirsch et al., 1998; Spalding et al., 1999), while evidence is accumulating that many K transporters, including members of the K transporter (KT)/high affinity K transporter (HAK)/K uptake permease (KUP) family, are low-affinity K transporters (Quintero and Blatt, 1997; Senn et al., 2001), implying that functions of plant K channels and transporters overlap at different K concentration ranges.Out of the three families of K transporters, cation proton antiporter (CPA), high affinity K/Na transporter (HKT), and KT/HAK/KUP, CPA was characterized as a K+(Na+)/H+ antiporter, HKT may cotransport Na and K or transport Na only (Rubio et al., 1995; Uozumi et al., 2000), while KT/HAK/KUP were predicted to be H+-coupled K+ symporters (Mäser et al., 2001; Lebaudy et al., 2007). KT/HAK/KUP were named by different researchers who first identified and cloned them (Quintero and Blatt, 1997; Santa-María et al., 1997). In plants, the KT/HAK/KUP family is the largest K transporter family, including 13 members in Arabidopsis and 27 members in the rice (Oryza sativa) genome (Rubio et al., 2000; Mäser et al., 2001; Bañuelos et al., 2002; Gupta et al., 2008). Sequence alignments show that genes of this family share relatively low homology to each other. The KT/HAK/KUP family was divided into four major clusters (Rubio et al., 2000; Gupta et al., 2008), and in cluster I and II, they were further separated into A and B groups. Genes of cluster I or II likely exist in all plants, cluster III is composed of genes from both Arabidopsis and rice, while cluster IV includes only four rice genes (Grabov, 2007; Gupta et al., 2008).The functions of KT/HAK/KUP were studied mostly in heterologous expression systems. Transporters of cluster I, such as AtHAK5, HvHAK1, OsHAK1, and OsHAK5, are localized in the plasma membrane (Kim et al., 1998; Bañuelos et al., 2002; Gierth et al., 2005) and exhibit high-affinity K uptake in the yeast Saccharomyces cerevisiae (Santa-María et al., 1997; Fu and Luan, 1998; Rubio et al., 2000) and in Escherichia coli (Horie et al., 2011). Transporters of cluster II, like AtKUP4 (TINY ROOT HAIRS1, TRH1), HvHAK2, OsHAK2, OsHAK7, and OsHAK10, could not complement the K uptake-deficient yeast (Saccharomyces cerevisiae) but were able to mediate K fluxes in a bacterial mutant; they might be tonoplast transporters (Senn et al., 2001; Bañuelos et al., 2002; Rodríguez-Navarro and Rubio, 2006). The function of transporters in clusters III and IV is even less known (Grabov, 2007).Existing data suggest that some KT/HAK/KUP transporters also may respond to salinity stress (Maathuis, 2009). The cluster I transporters of HvHAK1 mediate Na influx (Santa-María et al., 1997), while AtHAK5 expression is inhibited by Na (Rubio et al., 2000; Nieves-Cordones et al., 2010). Expression of OsHAK5 in tobacco (Nicotiana tabacum) BY2 cells enhanced the salt tolerance of these cells by accumulating more K without affecting their Na content (Horie et al., 2011).There are only scarce reports on the physiological function of KT/HAK/KUP in planta. In Arabidopsis, mutation of AtKUP2 (SHORT HYPOCOTYL3) resulted in a short hypocotyl, small leaves, and a short flowering stem (Elumalai et al., 2002), while a loss-of-function mutation of AtKUP4 (TRH1) resulted in short root hairs and a loss of gravity response in the root (Rigas et al., 2001; Desbrosses et al., 2003; Ahn et al., 2004). AtHAK5 is the only system currently known to mediate K uptake at concentrations below 0.01 mm (Rubio et al., 2010) and provides a cesium uptake pathway (Qi et al., 2008). AtHAK5 and AtAKT1 are the two major physiologically relevant molecular entities mediating K uptake into roots in the range between 0.01 and 0.05 mm (Pyo et al., 2010; Rubio et al., 2010). AtAKT1 may contribute to K uptake within the K concentrations that belong to the high-affinity system described by Epstein et al. (1963).Among all 27 members of the KT/HAK/KUP family in rice, OsHAK1, OsHAK5, OsHAK19, and OsHAK20 were grouped in cluster IB (Gupta et al., 2008). These four rice HAK members share 50.9% to 53.4% amino acid identity with AtHAK5. OsHAK1 was expressed in the whole plant, with maximum expression in roots, and was up-regulated by K deficiency; it mediated high-affinity K uptake in yeast (Bañuelos et al., 2002). In this study, we examined the tissue-specific localization and the physiological functions of OsHAK5 in response to variation in K supply and to salt stress in rice. By comparing K uptake and translocation in OsHAK5 knockout (KO) mutants and in OsHAK5-overexpressing lines with those in their respective wild-type lines supplied with different K concentrations, we found that OsHAK5 not only mediates high-affinity K acquisition but also participates in root-to-shoot K transport as well as in K-regulated salt tolerance.  相似文献   

12.
Ca2+-dependent protein kinases (CPKs) form a large family of 34 genes in Arabidopsis (Arabidopsis thaliana). Based on their dependence on Ca2+, CPKs can be sorted into three types: strictly Ca2+-dependent CPKs, Ca2+-stimulated CPKs (with a significant basal activity in the absence of Ca2+), and essentially calcium-insensitive CPKs. Here, we report on the third type of CPK, CPK13, which is expressed in guard cells but whose role is still unknown. We confirm the expression of CPK13 in Arabidopsis guard cells, and we show that its overexpression inhibits light-induced stomatal opening. We combine several approaches to identify a guard cell-expressed target. We provide evidence that CPK13 (1) specifically phosphorylates peptide arrays featuring Arabidopsis K+ Channel KAT2 and KAT1 polypeptides, (2) inhibits KAT2 and/or KAT1 when expressed in Xenopus laevis oocytes, and (3) closely interacts in plant cells with KAT2 channels (Förster resonance energy transfer-fluorescence lifetime imaging microscopy). We propose that CPK13 reduces stomatal aperture through its inhibition of the guard cell-expressed KAT2 and KAT1 channels.Stomata are microscopic organs at the leaf surface, each made of two so-called guard cells forming a pore. Opening or closing these pores is the way through which plants control their gas exchanges with the atmosphere (i.e. carbon dioxide uptake to feed the photosynthetic process and transpirational loss of water vapor). Stomatal movements result from osmotically driven fluxes of water, which follow massive exchanges of solutes, including K+ ions, between the guard cells and the surrounding tissues (Hetherington, 2001; Nilson and Assmann, 2007).Both Ca2+-dependent and Ca2+-independent signaling pathways are known to control stomatal movements (MacRobbie, 1993, 1998; Blatt, 2000; Webb et al., 2001; Mustilli et al., 2002; Israelsson et al., 2006; Marten et al., 2007; Laanemets et al., 2013). In particular, Ca2+ signals have been reported to promote stomatal closure through the inhibition of inward K+ channels and the activation of anion channels (Blatt, 1991, 1992, 2000; Thiel et al., 1992; Grabov and Blatt, 1999; Schroeder et al., 2001; Hetherington and Brownlee, 2004; Mori et al., 2006; Marten et al., 2007; Geiger et al., 2010; Brandt et al., 2012; Scherzer et al., 2012). However, little is known about the molecular identity of the links between Ca2+ events and Shaker K+ channel activity. Several kinases and phosphatases are believed to be involved in both the Ca2+-dependent and Ca2+-independent signaling pathways. Plants express two large kinase families whose activity is related to Ca2+ signaling. Firstly, CBL-interacting protein kinases (CIPKs; 25 genes in Arabidopsis [Arabidopsis thaliana]) are indirectly controlled by their interaction with a set of calcium sensors, the calcineurin B-like proteins (CBLs; 10 genes in Arabidopsis). This complex forms a fascinating network of potential Ca2+ signaling decoders (Luan, 2009; Weinl and Kudla, 2009), which have been addressed in numerous reports (Xu et al., 2006; Hu et al., 2009; Batistic et al., 2010; Held et al., 2011; Chen et al., 2013). In particular, some CBL-CIPK pairs have been shown to regulate Shaker channels such as Arabidopsis K+ Transporter1 (AKT1; Xu et al., 2006; Lan et al., 2011) or AKT2 (Held et al., 2011). Second, Ca2+-dependent protein kinases (CPKs) form an even larger family (34 genes in Arabidopsis) of proteins combining a kinase domain with the ability to bind Ca2+, thanks to the so-called EF hands (Harmon et al., 2000; Harper et al., 2004). CPKs, which, interestingly, are not found in animal cells, exhibit different calcium dependencies (Boudsocq et al., 2012). With respect to this, three types of CPKs can be considered: strictly Ca2+-dependent CPKs, Ca2+-stimulated CPKs (with a significant basal activity in the absence of Ca2+), and essentially Ca2+-insensitive CPKs (however, structurally close to kinases of groups 1 and 2).Pioneering work by Luan et al. (1993) demonstrated in Vicia faba guard cells that inward K+ channels were regulated by some Ca2+-dependent kinases. Then, such a Ca2+-dependent kinase was purified from guard cell protoplasts of V. faba and shown to actually phosphorylate the in vitro-translated KAT1 protein, a Shaker channel subunit natively expressed in Arabidopsis guard cells (Li et al., 1998). KAT1 regulation by CPK was shown by the inhibition of KAT1 currents after the coexpression of KAT1 and CDPK from soybean (Glycine max) in oocytes (Berkowitz et al., 2000). Since then, several cpk mutant lines of Arabidopsis have been shown to be impaired in stomatal movements, for example cpk10 (Ca2+ insensitive), cpk4/cpk11 (Ca2+ dependent), and cpk3/cpk6/cpk23 (Ca2+ dependent; Mori et al., 2006; Geiger et al., 2010; Munemasa et al., 2011; Hubbard et al., 2012).Of the nine genes encoding voltage-dependent K+ channels (Shaker) in Arabidopsis (Véry and Sentenac, 2002, 2003; Lebaudy et al., 2007; Hedrich, 2012), six are expressed in guard cells and play a role in stomatal movements: the Gated Outwardly-Rectifying K+ (GORK) gene, encoding an outward K+ channel subunit, and the AKT1, AKT2, Arabidopsis K+ Rectifying Channel1 (AtKC1), KAT1, and KAT2 genes, encoding inward K+ channel subunits (Pilot et al., 2001; Szyroki et al., 2001; Hosy et al., 2003; Pandey et al., 2007; Lebaudy et al., 2008a). Shaker channels result from the assembly of four subunits, and it has been shown that inward subunits tend to heterotetramerize, thus potentially widening the functional and regulatory scope of inward K+ conductance in guard cells (Xicluna et al., 2007; Jeanguenin et al., 2008; Lebaudy et al., 2008a, 2010). Inhibition of inward K+ channels has been shown to reduce stomatal opening (Liu et al., 2000; Kwak et al., 2001). This has grounded a strategy for disrupting inward K+ channel conductance in guard cells by expressing a nonfunctional KAT2 subunit (dominant negative mutation) in a kat2 knockout Arabidopsis line. The resulting Arabidopsis lines, named kincless, have no functional inward K+ channels and exhibit delayed stomatal opening (Lebaudy et al., 2008b) with, in the long term, a biomass reduction compared with the Arabidopsis wild-type line.Among the CPKs presumably expressed in Arabidopsis guard cells (Leonhardt et al., 2004), we looked for CPK13, which belongs to the atypical Ca2+-insensitive type of CPKs (Kanchiswamy et al., 2010; Boudsocq et al., 2012; Liese and Romeis, 2013) and whose role remains unknown in stomatal movements. Here, we confirm first that CPK13 kinase activity is independent of Ca2+ and show that CPK13 expression is predominant in Arabidopsis guard cells using CPK13-GUS lines. We then report that overexpression of CPK13 in Arabidopsis induces a dramatic default in stomatal aperture. Based on the previously reported kincless phenotype (Lebaudy et al., 2008b), we propose that CPK13 could reduce the activity of inward K+ channels in guard cells, particularly that of KAT2. We confirm this hypothesis by voltage-clamp experiments and show an inhibition of KAT2 and KAT1 activity by CPK13 (but not that of AKT2). In addition, we present peptide array phosphorylation assays showing that CPK13 targets, with some specificity, several KAT2 and KAT1 polypeptides. Finally, we demonstrate that KAT2 and CPK13 interact in planta using Förster resonance energy transfer (FRET)-fluorescence lifetime imaging microscopy (FLIM).  相似文献   

13.
14.
15.
16.
17.
Futile transmembrane NH3/NH4+ cycling in plant root cells, characterized by extremely rapid fluxes and high efflux to influx ratios, has been successfully linked to NH3/NH4+ toxicity. Surprisingly, the fundamental question of which species of the conjugate pair (NH3 or NH4+) participates in such fluxes is unresolved. Using flux analyses with the short-lived radioisotope 13N and electrophysiological, respiratory, and histochemical measurements, we show that futile cycling in roots of barley (Hordeum vulgare) seedlings is predominately of the gaseous NH3 species, rather than the NH4+ ion. Influx of 13NH3/13NH4+, which exceeded 200 µmol g–1 h–1, was not commensurate with membrane depolarization or increases in root respiration, suggesting electroneutral NH3 transport. Influx followed Michaelis-Menten kinetics for NH3 (but not NH4+), as a function of external concentration (Km = 152 µm, Vmax = 205 µmol g–1 h–1). Efflux of 13NH3/13NH4+ responded with a nearly identical Km. Pharmacological characterization of influx and efflux suggests mediation by aquaporins. Our study fundamentally revises the futile-cycling model by demonstrating that NH3 is the major permeating species across both plasmalemma and tonoplast of root cells under toxicity conditions.Ammonia/ammonium (NH3/NH4+) toxicity in higher plants has resulted in crop reduction and forest decline (Pearson and Stewart, 1993; Vitousek et al., 1997; Britto and Kronzucker, 2002), biodiversity loss (Stevens et al., 2004; Bobbink et al., 2010), and species extirpation (de Graaf et al., 1998; McClean et al., 2011). These major ecological and economic problems have been aggravated by an accelerated global nitrogen (N) cycle caused primarily by the industrialized production and use of N fertilizers (Galloway et al., 2008; Gruber and Galloway, 2008). With increasing global population and demands on agricultural production, there is no sign of this trend easing: anthropogenic N fixation has reached 210 teragrams year–1, an approximately 12% increase from 2005 and an approximately 1,300% rise from 150 years ago (Galloway et al., 2008; Fowler et al., 2013).Although considerable knowledge of the causes and mechanisms of NH3/NH4+ toxicity has accrued in recent years, our understanding of the key processes remains rudimentary (Gerendas et al., 1997; Britto and Kronzucker, 2002). A major hypothesis is that of futile transmembrane NH4+ cycling, which proposes a pathological inability of root cells to restrict the primary entry of NH4+ at high external concentrations ([NH4+]ext); many downstream toxicological events are contingent upon this entry (Britto et al., 2001b). In this model, a rapid, thermodynamically passive influx of NH4+ is coupled to an active efflux of NH4+ that is nearly as rapid, constraining normal cellular function and energetics and resulting in plant growth decline and mortality. This phenomenon is thought to occur in NH4+-sensitive species such as barley (Hordeum vulgare) and, to a lesser extent, in tolerant species such as rice (Oryza sativa), which can be susceptible at higher thresholds (Balkos et al., 2010; Chen et al., 2013).Most soils are typically acidic, especially when [NH4+] is high (i.e. in the millimolar range; Van Breemen et al., 1982; Bobbink et al., 1998; Britto and Kronzucker, 2002), and given the pKa of 9.25 for the conjugate pair NH3/NH4+, [NH3] is generally low (Izaurralde et al., 1990; Weise et al., 2013). Consequently, the fluxes of NH3 have largely been considered negligible (Britto et al., 2001a; Britto and Kronzucker, 2002; Loqué and von Wirén, 2004), in contrast to NH4+ fluxes, which are well characterized physiologically (Lee and Ayling, 1993; Wang et al., 1993a, 1993b; Kronzucker et al., 1996) and at the molecular level (Rawat et al., 1999; von Wirén et al., 2000; Ludewig et al., 2007), at least at lower concentrations. However, the transport of NH3 across membranes has received new attention in the light of evidence that some members of the aquaporin (AQP) family of transporters, a diverse and ubiquitous class of major intrinsic proteins (Maurel et al., 2008; Hove and Bhave, 2011), can mediate NH3 fluxes in single-cell systems (Jahn et al., 2004; Holm et al., 2005; Loqué et al., 2005; Saparov et al., 2007). However, a convincing demonstration that AQPs transport NH3 in planta is currently lacking. Given the unusually high capacity of AQP-mediated fluxes relative to those of ion channels and other transporters (Kozono et al., 2002), it is possible that sizable NH3 fluxes can be conducted through AQPs, even at very low external NH3 concentration ([NH3]ext).Here, we have critically reexamined the hypothesis that futile cycling is composed of cationic NH4+ fluxes across the plasmalemma, of which an active efflux mechanism accounts for energetic demands directly contributing to toxicity (Britto et al., 2001b). We present evidence for the following alternative scenario: 1) futile cycling consists mainly of the passive electroneutral flux of the conjugate base NH3; 2) such fluxes rapidly span both major membrane systems in root cells (i.e. plasmalemma and tonoplast); 3) AQPs mediate such fluxes; and 4) a thermodynamic equilibrium of NH3 is established throughout the cell, resulting in hyperaccumulation of NH4+ in the acidic vacuole. This evidence comes primarily from positron emission tracing with the short-lived radioisotope 13N, used to characterize the component fluxes of futile cycling at the cellular level in the model species barley. We have coupled this with 42K+ radiotracing, to provide comparison with a well-understood cationic flux, as well as electrophysiological, respiratory, pharmacological, and histochemical analyses.  相似文献   

18.
19.
20.
Salinity affects a significant portion of arable land and is particularly detrimental for irrigated agriculture, which provides one-third of the global food supply. Rice (Oryza sativa), the most important food crop, is salt sensitive. The genetic resources for salt tolerance in rice germplasm exist but are underutilized due to the difficulty in capturing the dynamic nature of physiological responses to salt stress. The genetic basis of these physiological responses is predicted to be polygenic. In an effort to address this challenge, we generated temporal imaging data from 378 diverse rice genotypes across 14 d of 90 mm NaCl stress and developed a statistical model to assess the genetic architecture of dynamic salinity-induced growth responses in rice germplasm. A genomic region on chromosome 3 was strongly associated with the early growth response and was captured using visible range imaging. Fluorescence imaging identified four genomic regions linked to salinity-induced fluorescence responses. A region on chromosome 1 regulates both the fluorescence shift indicative of the longer term ionic stress and the early growth rate decline during salinity stress. We present, to our knowledge, a new approach to capture the dynamic plant responses to its environment and elucidate the genetic basis of these responses using a longitudinal genome-wide association model.Nearly one-third of the 54 million ha of the highly saline soils in the world are located in South and Southeast Asia. Rice (Oryza sativa), which is the primary source of calories and protein for these two regions, is very sensitive to salinity stress, with even moderate salinity levels known to decrease yields by 50% (Zeng et al., 2002). Projected sea level rise due to climate change is expected to increase saltwater ingress in coastal rice-growing regions of South and Southeast Asia. Therefore, development of salt-tolerant rice cultivars is essential to maintain rice productivity in the salinity-affected regions globally.Salt tolerance, defined as the ability to maintain growth and productivity in saline conditions, is a complex polygenic trait that may be influenced by distinct physiological mechanisms (Munns et al., 1982; Munns and Termaat, 1986; Cheeseman, 1988; Munns and Tester, 2008; Horie et al., 2012; for a comprehensive review of genes involved in salinity tolerance in rice, see Negrão et al., 2011) At the cellular level, plants respond to saline conditions in two phases, namely an osmotic (shoot ion independent) and an ionic stress phase, which can occur in an overlapping manner with varying intensity during the course of salinity stress (Munns and Termaat, 1986; Munns, 2002; Munns and James, 2003; Munns and Tester, 2008; Horie et al., 2012). During the osmotic stress phase, which occurs soon after the onset of salinity, the reduction in external osmotic potential disrupts water uptake and impedes cell expansion, which, at the whole plant level, leads to reduced growth rate (Matsuda and Riazi, 1981; Munns and Passioura, 1984; Rawson and Munns, 1984; Azaizeh and Steudle, 1991; Fricke and Peters, 2002; Fricke, 2004; Boursiac et al., 2005). As salinity stress persists over several days and weeks, sodium ions (Na+) accumulate to toxic levels, resulting in cell death and precocious leaf senescence (Lutts and Bouharmont, 1996; Munns, 2002; Munns and James, 2003; Ghanem et al., 2008). This is typically observed during the ionic phase of the salinity response (Munns, 2002; Munns and James, 2003; Munns and Tester, 2008). Plants possess distinct mechanisms to adapt to these osmotic and ionic stresses that are controlled by a suite of genes (Apse et al., 1999; Carvajal et al., 1999; Halfter et al., 2000; Ishitani et al., 2000; Shi et al., 2000; Zeng and Shannon, 2000; Rus et al., 2001; Berthomieu et al., 2003; Martínez-Ballesta et al., 2003; Boursiac et al., 2005, 2008; Ren et al., 2005; Huang et al., 2006; Davenport et al., 2007; Obata et al., 2007; Székely et al., 2008; Horie et al., 2011; Rivandi et al., 2011; Asano et al., 2012; Munns et al., 2012; Latz et al., 2013; Schmidt et al., 2013; Campo et al., 2014; Choi et al., 2014; Liu et al., 2014). The genetic basis of temporal adaptive responses to salinity stress remains to be explored in rice and other crops. This is primarily due to challenges in capturing the dynamic physiological responses to salinity for a large number of genotypes in a nondestructive manner. Manual phenotyping to detect incremental changes in growth rate during the osmotic stress and slight shifts in leaf color due to ionic stress is difficult to quantify for a large number of genotypes.In rice, at least one major quantitative trait loci (QTL; saltol) for salinity tolerance has been characterized based on end point measurements of biomass, senescence/injury, and Na+ and K+ concentrations (Bonilla et al., 2002; Lin et al., 2004; Thomson et al., 2010). SHOOT K+ CONTENT1 (SKC1) is the causative gene underlying the saltol region. SKC1 encodes a Na+-selective high-affinity potassium transporter that regulates Na+/K+ homeostasis during salinity stress (Ren et al., 2005). High levels of Na+ displace cellular K+, an essential element for several enzymatic reactions and physiological processes (Gierth and Mäser, 2007). The ability to maintain cellular K+ levels during salinity through the action of Na+-selective potassium transporters or Na+/H+ antiporters is a well-characterized tolerance mechanism in cereals including rice (Ren et al., 2005; Sunarpi et al., 2005; Huang et al., 2006; Møller et al., 2009; Mian et al., 2011; Munns et al., 2012).Numerous studies have utilized conventional linkage mapping to identify QTL for morphological and physiological responses to salinity in rice using discrete end point measurements (Bonilla et al., 2002; Lin et al., 2004; Ren et al., 2005; Negrão et al., 2011; Wang et al., 2012). However, the physiological adaptation to saline conditions is a complex continuous process that develops over time. While some accessions will exhibit similar end point phenotypic values, the genetic and physiological mechanisms giving rise to the similar phenotypes may be very different and the growth trajectories throughout the experiment may be distinct. Although single time point studies have yielded important information regarding the genetic basis of salinity tolerance, such approaches are too simple to reveal the genetic architecture of stress adaptation. With the advent of high-throughput image-based phenotyping platforms, it is now feasible to quantify dynamic responses during the stress treatment for a large number of genotypes (Berger et al., 2010; Golzarian et al., 2011; Chen et al., 2014; Honsdorf et al., 2014).Image-based phenotyping has been combined with genome-wide association studies (GWAS) and linkage mapping to examine the genetic basis of complex developmental processes (Busemeyer et al., 2013; Moore et al., 2013; Topp et al., 2013; Slovak et al., 2014; Würschum et al., 2014; Yang et al., 2014; Bac-Molenaar et al., 2015). Moreover, the introduction of the time axis provides a better understanding of the physiological processes underlying complex stress and developmental responses compared with single end point measurements (Zhang et al., 2012; Moore et al., 2013; Brown et al., 2014; Chen et al., 2014; Slovak et al., 2014; Bac-Molenaar et al., 2015). However, to date, no studies have implemented an association mapping approach using image-derived phenotypes to address the genetic basis of dynamic stress responses in plants. Image-based phenotyping offers several advantages over conventional phenotyping: (1) quantitative measurements can be recorded over discrete time points to capture morphological and physiological responses in a nondestructive manner, and (2) the use of various types of spectral imaging address phenotypes that are not detectable to the human eye such as chlorophyll fluorescence and leaf water content. Integrating dynamic phenotypic data and association mapping has the potential to query genetic diversity across hundreds of accessions for complex traits and provides much higher resolution compared with conventional linkage mapping. Here, we explored the dynamic growth and chlorophyll responses to salinity of a diverse set of rice accessions using high-throughput visible and fluorescence imaging. To assess the genetic basis of plant growth in saline conditions, a logistic model was used to describe the temporal growth responses and was incorporated into the statistical framework necessary for association mapping. Coupled with temporal fluorescence imaging, we present, to our knowledge, new insights into the genetic architecture of osmotic and ionic responses during salinity stress in rice.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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