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.
Divinyl reductase (DVR) converts 8-vinyl groups on various chlorophyll intermediates to ethyl groups, which is indispensable for chlorophyll biosynthesis. To date, five DVR activities have been detected, but adequate evidence of enzymatic assays using purified or recombinant DVR proteins has not been demonstrated, and it is unclear whether one or multiple enzymes catalyze these activities. In this study, we systematically carried out enzymatic assays using four recombinant DVR proteins and five divinyl substrates and then investigated the in vivo accumulation of various chlorophyll intermediates in rice (Oryza sativa), maize (Zea mays), and cucumber (Cucumis sativus). The results demonstrated that both rice and maize DVR proteins can convert all of the five divinyl substrates to corresponding monovinyl compounds, while both cucumber and Arabidopsis (Arabidopsis thaliana) DVR proteins can convert three of them. Meanwhile, the OsDVR (Os03g22780)-inactivated 824ys mutant of rice exclusively accumulated divinyl chlorophylls in its various organs during different developmental stages. Collectively, we conclude that a single DVR with broad substrate specificity is responsible for reducing the 8-vinyl groups of various chlorophyll intermediates in higher plants, but DVR proteins from different species have diverse and differing substrate preferences, although they are homologous.Chlorophyll (Chl) molecules universally exist in photosynthetic organisms. As the main component of the photosynthetic pigments, Chl molecules perform essential processes of absorbing light and transferring the light energy in the reaction center of the photosystems (Fromme et al., 2003). Based on the number of vinyl side chains, Chls are classified into two groups, 3,8-divinyl (DV)-Chl and 3-monovinyl (MV)-Chl. The DV-Chl molecule contains two vinyl groups at positions 3 and 8 of the tetrapyrrole macrocycle, whereas the MV-Chl molecule contains a vinyl group at position 3 and an ethyl group at position 8 of the macrocycle. Almost all of the oxygenic photosynthetic organisms contain MV-Chls, with the exceptions of some marine picophytoplankton species that contain only DV-Chls as their primary photosynthetic pigments (Chisholm et al., 1992; Goericke and Repeta, 1992; Porra, 1997).The classical single-branched Chl biosynthetic pathway proposed by Granick (1950) and modified by Jones (1963) assumed the rapid reduction of the 8-vinyl group of DV-protochlorophyllide (Pchlide) catalyzed by a putative 8-vinyl reductase. Ellsworth and Aronoff (1969) found evidence for both MV and DV forms of several Chl biosynthetic intermediates between magnesium-protoporphyrin IX monomethyl ester (MPE) and Pchlide in Chlorella spp. mutants. Belanger and Rebeiz (1979, 1980) reported that the Pchlide pool of etiolated higher plants contains both MV- and DV-Pchlide. Afterward, following the further detection of MV- and DV-tetrapyrrole intermediates and their biosynthetic interconversion in tissues and extracts of different plants (Belanger and Rebeiz, 1982; Duggan and Rebeiz, 1982; Tripathy and Rebeiz, 1986, 1988; Parham and Rebeiz, 1992, 1995; Kim and Rebeiz, 1996), a multibranched Chl biosynthetic heterogeneity was proposed (Rebeiz et al., 1983, 1986, 1999; Whyte and Griffiths, 1993; Kolossov and Rebeiz, 2010).Biosynthetic heterogeneity refers to the biosynthesis of a particular metabolite by an organelle, tissue, or organism via multiple biosynthetic routes. Varieties of reports lead to the assumption that Chl biosynthetic heterogeneity originates mainly in parallel DV- and MV-Chl biosynthetic routes. These routes are interconnected by 8-vinyl reductases that convert DV-tetrapyrroles to MV-tetrapyrroles by conversion of the vinyl group at position 8 of ring B to the ethyl group (Parham and Rebeiz, 1995; Rebeiz et al., 2003). DV-MPE could be converted to MV-MPE in crude homogenates from etiolated wheat (Triticum aestivum) seedlings (Ellsworth and Hsing, 1974). Exogenous DV-Pchlide could be partially converted to MV-Pchlide in barley (Hordeum vulgare) plastids (Tripathy and Rebeiz, 1988). 8-Vinyl chlorophyllide (Chlide) a reductases in etioplast membranes isolated from etiolated cucumber (Cucumis sativus) cotyledons and barley and maize (Zea mays) leaves were found to be very active in the conversion of exogenous DV-Chlide a to MV-Chlide a (Parham and Rebeiz, 1992, 1995). Kim and Rebeiz (1996) suggested that Chl biosynthetic heterogeneity in higher plants may originate at the level of DV magnesium-protoporphyrin IX (Mg-Proto) and would be mediated by the activity of a putative 8-vinyl Mg-Proto reductase in barley etiochloroplasts and plastid membranes. However, since these reports did not use purified or recombinant enzyme, it is not clear whether the reductions of the 8-vinyl groups of various Chl intermediates are catalyzed by one enzyme of broad specificity or by multiple enzymes of narrow specificity, which actually has become one of the focus issues in Chl biosynthesis.Nagata et al. (2005) and Nakanishi et al. (2005) independently identified the AT5G18660 gene of Arabidopsis (Arabidopsis thaliana) as an 8-vinyl reductase, namely, divinyl reductase (DVR). Chew and Bryant (2007) identified the DVR BciA (CT1063) gene of the green sulfur bacterium Chlorobium tepidum, which is homologous to AT5G18660. An enzymatic assay using a recombinant Arabidopsis DVR (AtDVR) on five DV substrates revealed that the major substrate of AtDVR is DV-Chlide a, while the other four DV substrates could not be converted to corresponding MV compounds (Nagata et al., 2007). Nevertheless, a recombinant BciA is able to reduce the 8-vinyl group of DV-Pchlide to generate MV-Pchlide (Chew and Bryant, 2007). Recently, we identified the rice (Oryza sativa) DVR encoded by Os03g22780 that has sequence similarity with the Arabidopsis DVR gene AT5G18660. We also confirmed that the recombinant rice DVR (OsDVR) is able to not only convert DV-Chlide a to MV-Chlide a but also to convert DV-Chl a to MV-Chl a (Wang et al., 2010). Thus, it is possible that the reductions of the 8-vinyl groups of various Chl biosynthetic intermediates are catalyzed by one enzyme of broad specificity.In this report, we extended our studies to four DVR proteins and five DV substrates. First, ZmDVR and CsDVR genes were isolated from maize and cucumber genomes, respectively, using a homology-based cloning approach. Second, enzymatic assays were systematically carried out using recombinant OsDVR, ZmDVR, CsDVR, and AtDVR as representative DVR proteins and using DV-Chl a, DV-Chlide a, DV-Pchlide a, DV-MPE, and DV-Mg-Proto as DV substrates. Third, we examined the in vivo accumulations of various Chl intermediates in rice, maize, and cucumber. Finally, we systematically investigated the in vivo accumulations of Chl and its various intermediates in the OsDVR (Os03g22780)-inactivated 824ys mutant of rice (Wang et al., 2010). The results strongly suggested that a single DVR protein with broad substrate specificity is responsible for reducing the 8-vinyl groups of various intermediate molecules of Chl biosynthesis in higher plants, but DVR proteins from different species could have diverse and differing substrate preferences even though they are homologous.  相似文献   

2.
3.
Eucalyptus camaldulensis is a tree species in the Myrtaceae that exhibits extremely high resistance to aluminum (Al). To explore a novel mechanism of Al resistance in plants, we examined the Al-binding ligands in roots and their role in Al resistance of E. camaldulensis. We identified a novel type of Al-binding ligand, oenothein B, which is a dimeric hydrolyzable tannin with many adjacent phenolic hydroxyl groups. Oenothein B was isolated from root extracts of E. camaldulensis by reverse-phase high-performance liquid chromatography and identified by nuclear magnetic resonance and mass spectrometry analyses. Oenothein B formed water-soluble or -insoluble complexes with Al depending on the ratio of oenothein B to Al and could bind at least four Al ions per molecule. In a bioassay using Arabidopsis (Arabidopsis thaliana), Al-induced inhibition of root elongation was completely alleviated by treatment with exogenous oenothein B, which indicated the capability of oenothein B to detoxify Al. In roots of E. camaldulensis, Al exposure enhanced the accumulation of oenothein B, especially in EDTA-extractable forms, which likely formed complexes with Al. Oenothein B was localized mostly in the root symplast, in which a considerable amount of Al accumulated. In contrast, oenothein B was not detected in three Al-sensitive species, comprising the Myrtaceae tree Melaleuca bracteata, Populus nigra, and Arabidopsis. Oenothein B content in roots of five tree species was correlated with their Al resistance. Taken together, these results suggest that internal detoxification of Al by the formation of complexes with oenothein B in roots likely contributes to the high Al resistance of E. camaldulensis.Aluminum (Al) toxicity is a major factor that limits plant growth in acid soils and affects approximately 30% of the total ice-free land area of the world (von Uexküll and Mutert, 1995). Although Al in soils exist in nonphytotoxic silicate or oxide forms at neutral pH, it is solubilized into a phytotoxic form, mainly as Al3+, at a pH of less than 5 (Kinraide, 1991; Kochian, 1995). The accumulation of Al in root tips causes rapid inhibition of root elongation, which is a characteristic symptom of Al toxicity in plants (Delhaize and Ryan, 1995; Ma, 2007). In general, plants exhibit an inhibition of root elongation as early as 30 to 120 min after exposure to excessive Al (Barceló and Poschenrieder, 2002). Inhibition of root elongation leads to decreased water and nutrient uptake and, eventually, to restriction of growth of the whole plant.Plants have evolved different levels of Al resistance mediated by two distinct classes of mechanisms (Kochian et al., 2004; Ma, 2007). One strategy is the exclusion of Al from the root tips (exclusion mechanism), and the other is tolerance to Al that enters the root tips (internal tolerance mechanism). The secretion of organic acid anions from roots in response to exposure to Al is the best-documented mechanism for Al exclusion. Organic acid anions (i.e. malate, citrate, and oxalate) can form a complex with Al in the rhizosphere and thereby prevent Al from entering the root tips. The genes encoding transporters for the Al-induced secretion of malate and citrate have been identified and characterized in several plant species (Ryan et al., 2011; Delhaize et al., 2012). Organic acid anions also play a role in the detoxification of Al that enters the roots by means of internal formation of complexes with Al (Ma et al., 1998). However, findings in recent studies increasingly suggest that the Al resistance of some plant species and cultivars cannot be explained solely by these two functions of organic acid anions (Wenzl et al., 2001, 2002; Piñeros et al., 2005; Zheng et al., 2005; Famoso et al., 2010). In addition to organic acid anions, flavonoid-type phenolics (Kidd et al., 2001), phenolic compounds (Ofei-Manu et al., 2001), cyclic hydroxamates (Poschenrieder et al., 2005), and proanthocyanidins (Osawa et al., 2011) in roots or root exudates are proposed as potential organic ligands for Al. The mechanisms by which these additional ligands confer Al resistance remain poorly understood.Eucalyptus camaldulensis is an evergreen tree belonging to the Myrtaceae family and is cultivated in tropical and subtropical regions of the world on account of its superior growth, broad adaptability, and multipurpose wood properties. E. camaldulensis can grow in acid soils and even in acid sulfate soils, where the pH is often lower than 3.5 and the Al concentration in the soil solution often reaches the millimolar level (van Breemen and Pons, 1978). Indeed, seedlings of this species show no inhibition of root elongation and plant growth when exposed to 1 mm Al for 20 d under hydroponic conditions (Tahara et al., 2005). Such Al resistance is considerably higher than that reported for a variety of herbaceous crops and model plants in studies of Al resistance mechanisms; such plants exhibit an inhibition of root elongation at 1 to 50 μm Al (Wenzl et al., 2001). Although our understanding of Al resistance mechanisms in some crops and model plants has improved recently, that for extremely Al-resistant species such as E. camaldulensis is limited.In E. camaldulensis, citrate secretion from roots and its content in the root tips are increased by exposure to Al, suggesting that citrate may contribute to its Al resistance (Tahara et al., 2008a). However, the amounts of organic acid anions, including citrate, secreted from roots and contained within the root tips are lower than those of more sensitive species (Tahara et al., 2008a). Therefore, the high Al resistance of E. camaldulensis cannot be explained only by the presence of organic acid anions. Roots of E. camaldulensis can accumulate large amounts of Al (11 mg g−1 dry weight) with no symptoms of Al toxicity (Tahara et al., 2005), suggesting the existence of additional mechanisms for internal tolerance. In this study, we investigated the presence of novel Al-binding ligands other than organic acid anions in E. camaldulensis roots and identified a hydrolyzable tannin, oenothein B, as a novel type of Al-binding ligand. We also examined the role of the ligand in the internal Al tolerance of E. camaldulensis.  相似文献   

4.
The medicinal plant Salvia miltiorrhiza produces various tanshinone diterpenoids that have pharmacological activities such as vasorelaxation against ischemia reperfusion injury and antiarrhythmic effects. Their biosynthesis is initiated from the general diterpenoid precursor (E,E,E)-geranylgeranyl diphosphate by sequential reactions catalyzed by copalyl diphosphate synthase (CPS) and kaurene synthase-like cyclases. Here, we report characterization of these enzymatic families from S. miltiorrhiza, which has led to the identification of unique pathways, including roles for separate CPSs in tanshinone production in roots versus aerial tissues (SmCPS1 and SmCPS2, respectively) as well as the unique production of ent-13-epi-manoyl oxide by SmCPS4 and S. miltiorrhiza kaurene synthase-like2 in floral sepals. The conserved SmCPS5 is involved in gibberellin plant hormone biosynthesis. Down-regulation of SmCPS1 by RNA interference resulted in substantial reduction of tanshinones, and metabolomics analysis revealed 21 potential intermediates, indicating a complex network for tanshinone metabolism defined by certain key biosynthetic steps. Notably, the correlation between conservation pattern and stereochemical product outcome of the CPSs observed here suggests a degree of correlation that, especially when combined with the identity of certain key residues, may be predictive. Accordingly, this study provides molecular insights into the evolutionary diversification of functional diterpenoids in plants.Salvia miltiorrhiza, a Lamiaceae species known as red sage or tanshen, is a traditional Chinese medicinal herb that is described in Shen Nong Ben Cao Jing, the oldest classical Chinese herbal book, which dates from between 25 and 220 C.E. The lipophilic pigments from the reddish root and rhizome consist of abietane quinone diterpenoids (Nakao and Fukushima, 1934), largely tanshinone IIA, cryptotanshinone, and tanshinone I (Zhong et al., 2009). These are highly bioactive. For example, tanshinone IIA exerts vasorelaxative activity, has antiarrhythmic effects, provides protection against ischemia reperfusion injury (Zhou et al., 2005; Gao et al., 2008; Sun et al., 2008), and exhibits anticancer activities (Efferth et al., 2008; Lee et al., 2008; Wang et al., 2008; Gong et al., 2011). In addition, tanshinones have been reported to have a broad spectrum of antimicrobial activities against various plant pathogens, including rice (Oryza sativa) blast fungus Magnaporthe oryzae (Zhao et al., 2011). Although tanshinones are mainly accumulated in the roots, trace amounts of tanshinones have been detected in aerial organs as well (Hang et al., 2008).Diterpenoid biosynthesis is initiated by diterpene synthases (diTPSs), which catalyze cyclization and/or rearrangement of the general acyclic precursor (E,E,E)-geranylgeranyl diphosphate (GGPP) to form various hydrocarbon backbone structures that are precursors to more specific families of diterpenoids (Zi et al., 2014). Previous work has indicated that tanshinone biosynthesis is initiated by cyclization of GGPP to copalyl diphosphate (CPP) by a CPP synthase (SmCPS1) and subsequent further cyclization to the abietane miltiradiene by a kaurene synthase-like cyclase (SmKSL1), so named for its homology to the ent-kaurene synthases (KSs) required for GA plant hormone biosynthesis (Gao et al., 2009). Miltiradiene is a precursor to at least cryptotanshinone (Guo et al., 2013), and RNA interference (RNAi) knockdown of SmCPS1 expression reduces tanshinone production, at least in hairy root cultures (Cheng et al., 2014). The identification of SmCPS1 and SmKSL1 has been followed by that of many related diTPSs from other Lamiaceae plant species (Caniard et al., 2012; Sallaud et al., 2012; Schalk et al., 2012; Brückner et al., 2014; Pateraki et al., 2014). These largely exhibit analogous activity, particularly the CPSs, which produce CPP or the stereochemically related 8α-hydroxy-labd-13E-en-15-yl diphosphate (LDPP) rather than the enantiomeric (ent) CPP relevant to GA biosynthesis.To further investigate diterpenoid biosynthesis in S. miltiorrhiza, we report here a more thorough characterization of its diTPS family. A previously reported whole-genome shotgun sequencing survey (Ma et al., 2012) has indicated that there are at least five CPSs, although only two KSL genes in S. miltiorrhiza (Supplemental Table S1). Intriguingly, based on a combination of biochemical and genetic (RNAi gene silencing) evidence, we find that these diTPSs nevertheless account for at least four different diterpenoid biosynthetic pathways, each dependent on a unique CPS, with the KS presumably involved in GA biosynthesis seeming to be responsible for alternative diterpenoid metabolism as well. In addition, our studies clarify the evolutionary basis for the observed functional diversity, with investigation of gene structure, positive selection, molecular docking, and mutational analysis used to explore the driving force for the functional divergence of these diTPSs. Moreover, we report metabolomic analysis, also carried out with SmCPS1 RNAi lines, which enables prediction of the downstream steps in tanshinone biosynthesis.  相似文献   

5.
The multifunctional movement protein (MP) of Tomato mosaic tobamovirus (ToMV) is involved in viral cell-to-cell movement, symptom development, and resistance gene recognition. However, it remains to be elucidated how ToMV MP plays such diverse roles in plants. Here, we show that ToMV MP interacts with the Rubisco small subunit (RbCS) of Nicotiana benthamiana in vitro and in vivo. In susceptible N. benthamiana plants, silencing of NbRbCS enabled ToMV to induce necrosis in inoculated leaves, thus enhancing virus local infectivity. However, the development of systemic viral symptoms was delayed. In transgenic N. benthamiana plants harboring Tobacco mosaic virus resistance-22 (Tm-22), which mediates extreme resistance to ToMV, silencing of NbRbCS compromised Tm-22-dependent resistance. ToMV was able to establish efficient local infection but was not able to move systemically. These findings suggest that NbRbCS plays a vital role in tobamovirus movement and plant antiviral defenses.Plant viruses use at least one movement protein (MP) to facilitate viral spread between plant cells via plasmodesmata (PD; Lucas and Gilbertson, 1994; Ghoshroy et al., 1997). Among viral MPs, the MP of tobamoviruses, such as Tobacco mosaic virus (TMV) and its close relative Tomato mosaic virus (ToMV), is the best characterized. TMV MP specifically accumulates in PD and modifies the plasmodesmatal size exclusion limit in mature source leaves or tissues (Wolf et al., 1989; Deom et al., 1990; Ding et al., 1992). TMV MP and viral genomic RNA form a mobile ribonucleoprotein complex that is essential for cell-to-cell movement of viral infection (Watanabe et al., 1984; Deom et al., 1987; Citovsky et al., 1990, 1992; Kiselyova et al., 2001; Kawakami et al., 2004; Waigmann et al., 2007). TMV MP also enhances intercellular RNA silencing (Vogler et al., 2008) and affects viral symptom development, host range, and host susceptibility to virus (Dardick et al., 2000; Bazzini et al., 2007). Furthermore, ToMV MP is identified as an avirulence factor that is recognized by tomato (Solanum lycopersicum) resistance proteins Tobacco mosaic virus resistance-2 (Tm-2) and Tm-22 (Meshi et al., 1989; Lanfermeijer et al., 2004). Indeed, tomato Tm-22 confers extreme resistance against TMV and ToMV in tomato plants and even in heterologous tobacco (Nicotiana tabacum) plants (Lanfermeijer et al., 2003, 2004).To date, several host factors that interact with TMV MP have been identified. These TMV MP-binding host factors include cell wall-associated proteins such as pectin methylesterase (Chen et al., 2000), calreticulin (Meshi et al., 1989), ANK1 (Ueki et al., 2010), and the cellular DnaJ-like protein MPIP1 (Shimizu et al., 2009). Many cytoskeletal components such as actin filaments (McLean et al., 1995), microtubules (Heinlein et al., 1995), and the microtubule-associated proteins MPB2C (Kragler et al., 2003) and EB1a (Brandner et al., 2008) also interact with TMV MP. Most of these factors are involved in TMV cell-to-cell movement.Rubisco catalyzes the first step of CO2 assimilation in photosynthesis and photorespiration. The Rubisco holoenzyme is a heteropolymer consisting of eight large subunits (RbCLs) and eight small subunits (RbCSs). RbCL was reported to interact with the coat protein of Potato virus Y (Feki et al., 2005). Both RbCS and RbCL were reported to interact with the P3 proteins encoded by several potyviruses, including Shallot yellow stripe virus, Onion yellow dwarf virus, Soybean mosaic virus, and Turnip mosaic virus (Lin et al., 2011). Proteomic analysis of the plant-virus interactome revealed that RbCS participates in the formation of virus complexes of Rice yellow mottle virus (Brizard et al., 2006). However, the biological function of Rubisco in viral infection remains unknown.In this study, we show that RbCS plays an essential role in virus movement, host susceptibility, and Tm-22-mediated extreme resistance in the ToMV-host plant interaction.  相似文献   

6.
7.
8.
Many unicellular microalgae produce large amounts (∼20 to 50% of cell dry weight) of triacylglycerols (TAGs) under stress (e.g., nutrient starvation and high light), but the synthesis and physiological role of TAG are poorly understood. We present detailed genetic, biochemical, functional, and physiological analyses of phospholipid:diacylglycerol acyltransferase (PDAT) in the green microalga Chlamydomonas reinhardtii, which catalyzes TAG synthesis via two pathways: transacylation of diacylglycerol (DAG) with acyl groups from phospholipids and galactolipids and DAG:DAG transacylation. We demonstrate that PDAT also possesses acyl hydrolase activities using TAG, phospholipids, galactolipids, and cholesteryl esters as substrates. Artificial microRNA silencing of PDAT in C. reinhardtii alters the membrane lipid composition, reducing the maximum specific growth rate. The data suggest that PDAT-mediated membrane lipid turnover and TAG synthesis is essential for vigorous growth under favorable culture conditions and for membrane lipid degradation with concomitant production of TAG for survival under stress. The strong lipase activity of PDAT with broad substrate specificity suggests that this enzyme could be a potential biocatalyst for industrial lipid hydrolysis and conversion, particularly for biofuel production.  相似文献   

9.
A large portion of the volatile organic compounds emitted by plants are oxygenated to yield reactive carbonyl species, which have a big impact on atmospheric chemistry. Deposition to vegetation driven by the absorption of reactive carbonyl species into plants plays a major role in cleansing the atmosphere, but the mechanisms supporting this absorption have been little examined. Here, we performed model experiments using methacrolein (MACR), one of the major reactive carbonyl species formed from isoprene, and tomato (Solanum lycopersicum) plants. Tomato shoots enclosed in a jar with MACR vapor efficiently absorbed MACR. The absorption efficiency was much higher than expected from the gas/liquid partition coefficient of MACR, indicating that MACR was likely metabolized in leaf tissues. Isobutyraldehyde, isobutyl alcohol, and methallyl alcohol (MAA) were detected in the headspace and inside tomato tissues treated with MACR vapor, suggesting that MACR was enzymatically reduced. Glutathione (GSH) conjugates of MACR (MACR-GSH) and MAA (MAA-GSH) were also detected. MACR-GSH was essentially formed through spontaneous conjugation between endogenous GSH and exogenous MACR, and reduction of MACR-GSH to MAA-GSH was likely catalyzed by an NADPH-dependent enzyme in tomato leaves. Glutathionylation was the metabolic pathway most responsible for the absorption of MACR, but when the amount of MACR exceeded the available GSH, MACR that accumulated reduced photosynthetic capacity. In an experiment simulating the natural environment using gas flow, MACR-GSH and MAA-GSH accumulation accounted for 30% to 40% of the MACR supplied. These results suggest that MACR metabolism, especially spontaneous glutathionylation, is an essential factor supporting MACR absorption from the atmosphere by tomato plants.Plants emit vast amounts of volatile organic chemicals (VOCs) into the atmosphere. The annual emission of VOCs other than methane is estimated to be approximately 1,300 Tg of carbon (Goldstein and Galbally, 2007), with approximately 90% originating from biogenic sources, of which one-third (approximately 500 Tg of carbon/year) is isoprene (Guenther et al., 1995). In the atmosphere, VOCs undergo the chemical processes of photolysis and reaction with hydroxyl and nitrate radicals (Atkinson and Arey, 2003). Isoprene, for example, is converted into a series of isomeric hydroxyl-substituted alkyl peroxyl radicals, which are further converted into methyl vinyl ketone (MVK; but-3-en-2-one) and methacrolein (MACR; 2-methylprop-2-enal; Liu et al., 2013). These VOCs and their oxygenated products (oVOCs) are important components for the production of ozone and aerosols, and thus have a big impact on atmospheric chemistry and even on the climate system (Goldstein and Galbally, 2007). VOCs and oVOCs are removed from the atmosphere through oxidation to carbon monoxide or dioxide, dry or wet deposition, or secondary aerosol formation (Goldstein and Galbally, 2007). Among these, deposition to vegetation plays a major role in the removal of VOCs and oVOCs from the atmosphere (Karl et al., 2010).A significant portion of the deposition to vegetation is attributable to the uptake of VOCs and oVOCs by plants, and a field study showed that MVK and MACR were immediately lost once they entered a leaf through stomata (Karl et al., 2010). Under growth conditions where stomatal conductance is high enough, the partitioning of VOCs between air and leaf water phases in equilibrium and the capacity of the plant to metabolize, translocate, and store VOCs determine their uptake rate (Tani et al., 2013). The immediate loss in leaves observed with MVK and MACR is indicative of efficient enzymatic reactions metabolizing them; however, the details of the metabolism of these oVOCs have been little investigated so far.The absorption and metabolism of several VOCs by plants have been reported. Airborne ent-kaurene was absorbed by Arabidopsis (Arabidopsis thaliana), Japanese cypress (Chamaecyparis obtusa), and Japanese cedar (Cryptomeria japonica) plants and converted into GAs (Otsuka et al., 2004). Arabidopsis absorbed (Z)-3-hexenal and converted it into (Z)-3-hexen-1-ol or further into (Z)-3-hexen-1-yl acetate using NADPH and acetyl-CoA, probably inside the plant tissues (Matsui et al., 2012). Nicotiana attenuata plants absorbed dimethyl disulfide formed by rhizobacteria (Meldau et al., 2013). The sulfur atom derived from volatile dimethyl disulfide was assimilated into plant proteins. Karl et al. (2010) assumed that aldehyde dehydrogenase, which is involved in detoxification that limits aldehyde accumulation and oxidative stress (Kirch et al., 2004), is involved in the uptake of oVOCs containing an aldehyde moiety; however, they did not provide direct evidence supporting their assumption.Conjugation of VOCs and oVOCs with sugar or glutathione (GSH) is another way to metabolize them. (Z)-3-Hexen-1-ol in the vapor phase was taken up by tomato (Solanum lycopersicum) plants and converted into its glycoside (Sugimoto et al., 2014). (E)-2-Hexenal reacts with GSH spontaneously and/or via glutathione S-transferase (GST) to form hexanal-GSH, which is subsequently reduced to hexanol-GSH (Davoine et al., 2006), although it is uncertain whether airborne (E)-2-hexenal is converted into its corresponding GSH adduct. Glutathionylation of (E)-2-hexenal is common and has been confirmed in grapevine (Vitis vinifera) and passion fruit (Passiflora edulis; Kobayashi et al., 2011; Fedrizzi et al., 2012). The catabolites formed from the GSH adduct in these crops are precursors for important flavor components.Although it is clear that oVOCs are absorbed by vegetation and that their efficient uptake is probably supported by metabolism in plant tissues, the metabolic fates of oVOCs taken up from the vapor phase into plants have been little studied. Here, we performed a series of model experiments using tomato seedlings and MACR to dissect the fates of oVOCs once they entered into plant tissues. To clearly see absorption of MACR and its fates in plant tissues, a model experiment under enclosed conditions with a high concentration of MACR was first carried out. Subsequently, an airflow system with a realistically low concentration of MACR was used. Tomato plants efficiently absorbed MACR. Reduction of the carbonyl moiety and the double bond conjugated to the carbonyl and conjugation with GSH were the major methods of metabolism of exogenous MACR. The metabolism seemed to be involved in the detoxification of reactive carbonyl species, which, in turn, accounted for the oVOC deposition to vegetation.  相似文献   

10.
11.
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   

12.
13.
Plants differ greatly in their susceptibility to insect herbivory, suggesting both local adaptation and resistance tradeoffs. We used maize (Zea mays) recombinant inbred lines to map a quantitative trait locus (QTL) for the maize leaf aphid (Rhopalosiphum maidis) susceptibility to maize Chromosome 1. Phytochemical analysis revealed that the same locus was also associated with high levels of 2-hydroxy-4,7-dimethoxy-1,4-benzoxazin-3-one glucoside (HDMBOA-Glc) and low levels of 2,4-dihydroxy-7-methoxy-1,4-benzoxazin-3-one glucoside (DIMBOA-Glc). In vitro enzyme assays with candidate genes from the region of the QTL identified three O-methyltransferases (Bx10a-c) that convert DIMBOA-Glc to HDMBOA-Glc. Variation in HDMBOA-Glc production was attributed to a natural CACTA family transposon insertion that inactivates Bx10c in maize lines with low HDMBOA-Glc accumulation. When tested with a population of 26 diverse maize inbred lines, R. maidis produced more progeny on those with high HDMBOA-Glc and low DIMBOA-Glc. Although HDMBOA-Glc was more toxic to R. maidis than DIMBOA-Glc in vitro, BX10c activity and the resulting decline of DIMBOA-Glc upon methylation to HDMBOA-Glc were associated with reduced callose deposition as an aphid defense response in vivo. Thus, a natural transposon insertion appears to mediate an ecologically relevant trade-off between the direct toxicity and defense-inducing properties of maize benzoxazinoids.  相似文献   

14.
The signaling role of hydrogen gas (H2) has attracted increasing attention from animals to plants. However, the physiological significance and molecular mechanism of H2 in drought tolerance are still largely unexplored. In this article, we report that abscisic acid (ABA) induced stomatal closure in Arabidopsis (Arabidopsis thaliana) by triggering intracellular signaling events involving H2, reactive oxygen species (ROS), nitric oxide (NO), and the guard cell outward-rectifying K+ channel (GORK). ABA elicited a rapid and sustained H2 release and production in Arabidopsis. Exogenous hydrogen-rich water (HRW) effectively led to an increase of intracellular H2 production, a reduction in the stomatal aperture, and enhanced drought tolerance. Subsequent results revealed that HRW stimulated significant inductions of NO and ROS synthesis associated with stomatal closure in the wild type, which were individually abolished in the nitric reductase mutant nitrate reductase1/2 (nia1/2) or the NADPH oxidase-deficient mutant rbohF (for respiratory burst oxidase homolog). Furthermore, we demonstrate that the HRW-promoted NO generation is dependent on ROS production. The rbohF mutant had impaired NO synthesis and stomatal closure in response to HRW, while these changes were rescued by exogenous application of NO. In addition, both HRW and hydrogen peroxide failed to induce NO production or stomatal closure in the nia1/2 mutant, while HRW-promoted ROS accumulation was not impaired. In the GORK-null mutant, stomatal closure induced by ABA, HRW, NO, or hydrogen peroxide was partially suppressed. Together, these results define a main branch of H2-regulated stomatal movement involved in the ABA signaling cascade in which RbohF-dependent ROS and nitric reductase-associated NO production, and subsequent GORK activation, were causally involved.Stomata are responsible for leaves of terrestrial plants taking in carbon dioxide for photosynthesis and likewise regulate how much water plants evaporate through the stomatal pores (Chaerle et al., 2005). When experiencing water-deficient conditions, surviving plants balance photosynthesis with controlling water loss through the stomatal pores, which relies on turgor changes by pairs of highly differentiated epidermal cells surrounding the stomatal pore, called the guard cells (Haworth et al., 2011; Loutfy et al., 2012).Besides the characterization of the significant roles of abscisic acid (ABA) in regulating stomatal movement, the key factors in guard cell signal transduction have been intensively investigated by performing forward and reverse genetics approaches. For example, both reactive oxygen species (ROS) and nitric oxide (NO) have been identified as vital intermediates in guard cell ABA signaling (Bright et al., 2006; Yan et al., 2007; Suzuki et al., 2011; Hao et al., 2012). The key ROS-producing enzymes in Arabidopsis (Arabidopsis thaliana) guard cells are the respiratory burst oxidase homologs (Rboh) D and F (Kwak et al., 2003; Bright et al., 2006; Mazars et al., 2010; Marino et al., 2012). Current available data suggest that there are at least two distinct pathways responsible for NO synthesis involved in ABA signaling in guard cells: the nitrite reductase (NR)- and l-Arg-dependent pathways (Desikan et al., 2002; Besson-Bard et al., 2008). Genetic evidence further demonstrated that removal of the major known sources of either ROS or NO significantly impairs ABA-induced stomatal closure. ABA fails to induce ROS production in the atrbohD/F double mutant (Kwak et al., 2003; Wang et al., 2012) and NO synthesis in the NR-deficient mutant nitrate reductase1/2 (nia1/2; Bright et al., 2006; Neill et al., 2008), both of which lead to impaired stomatal closure in Arabidopsis. Most importantly, ROS and NO, which function both synergistically and independently, have been established as ubiquitous signal transduction components to control a diverse range of physiological pathways in higher plants (Bright et al., 2006; Tossi et al., 2012).The guard cell outward-rectifying K+ channel (GORK) encodes the exclusive voltage-gated outwardly rectifying K+ channel protein, which was located in the guard cell membrane (Ache et al., 2000; Dreyer and Blatt, 2009). Expression profiles revealed that this gene is up-regulated upon the onset of drought, salinity, and cold stress and ABA exposure (Becker et al., 2003; Tran et al., 2013). Reverse genetic evidence further showed that GORK plays an important role in the control of stomatal movements and allows the plant to reduce transpirational water loss significantly (Hosy et al., 2003) and participates in the regulation of salinity tolerance by preventing salt-induced K+ loss (Jayakannan et al., 2013). Due to the high complexity of guard cell signaling cascades, whether and how ABA-triggered GORK up-regulation is attributed to the generation of cellular secondary messengers, such as ROS and NO, is less clear.Hydrogen gas (H2) was recently revealed as a signaling modulator with multiple biological functions in clinical trails (Ohsawa et al., 2007; Itoh et al., 2009; Ito et al., 2012). It was previously found that a hydrogenase system could generate H2 in bacteria and green algae (Meyer, 2007; Esquível et al., 2011). Although some earlier studies discovered the evolution of H2 in several higher plant species (Renwick et al., 1964; Torres et al., 1984), it was also proposed that the eukaryotic hydrogenase-like protein does not metabolize H2 (Cavazza et al., 2008; Mondy et al., 2014). Since the explosion limit of H2 gas is about 4% to 72.4% (v/v, in the air), the direct application of H2 gas in experiments is flammable and dangerous. Regardless of these problems to be resolved, the methodology, such as using exogenous hydrogen-rich water (HRW) or hydrogen-rich saline, which is safe, economical, and easily available, provides a valuable approach to investigate the physiological function of H2 in animal research and clinical trials. For example, hydrogen dissolved in Dulbecco’s modified Eagle’s medium was found to react with cytotoxic ROS and thus protect against oxidative damage in PC12 cells and rats (Ohsawa et al., 2007). The neuroprotective effect of H2-loaded eye drops on retinal ischemia-reperfusion injury was also reported (Oharazawa et al., 2010). In plants, corresponding results by using HRW combined with gas chromatography (GC) revealed that H2 could act as a novel beneficial gaseous molecule in plant responses against salinity (Xie et al., 2012; Xu et al., 2013), cadmium stress (Cui et al., 2013), and paraquat toxicity (Jin et al., 2013). More recently, the observation that HRW could delay the postharvest ripening and senescence of kiwifruit (Actinidia deliciosa) was reported (Hu et al., 2014).Considering the fact that the signaling cascades for salt, osmotic, and drought stresses share a common cascade in an ABA-dependent pathway, it would be noteworthy to identify whether and how H2 regulates the bioactivity of ABA-induced downstream components and, thereafter, biological responses, including stomatal closure and drought tolerance. To resolve these scientific questions, rbohD, rbohF, nia1/2, nitric oxide associated1 (noa1; Van Ree et al., 2011), nia1/2/noa1, and gork mutants were utilized to investigate the relationship among H2, ROS, NO, and GORK in the guard cell signal transduction network. By the combination of pharmacological and biochemical analyses with this genetics-based approach, we provide comprehensive evidence to show that H2 might be a newly identified bioeffective modulator involved in ABA signaling responsible for drought tolerance, that HRW-promoted stomatal closure was mainly attributed to the modulation of ROS-dependent NO generation, and that GORK might be the downstream target protein of H2 signaling.  相似文献   

15.
Cytokinins (CKs), a class of phytohormones that regulate plant growth and development, are also synthesized by some phytopathogens to disrupt the hormonal balance and to facilitate niche establishment in their hosts. Rhodococcus fascians harbors the fasciation (fas) locus, an operon encoding several genes homologous to CK biosynthesis and metabolism. This pathogen causes unique leafy gall symptoms reminiscent of CK overproduction; however, bacterial CKs have not been clearly correlated with the severe symptoms, and no virulence-associated unique CKs or analogs have been identified. Here, we report the identification of monomethylated N6-(∆2-isopentenyl)adenine and dimethylated N6-(∆2-isopentenyl)adenine (collectively, methylated cytokinins [MeCKs]) from R. fascians. MeCKs were recognized by a CK receptor and up-regulated type-A ARABIDOPSIS THALIANA RESPONSE REGULATOR genes. Treatment with MeCKs inhibited root growth, a hallmark of CK action, whereas the receptor mutant was insensitive. MeCKs were retained longer in planta than canonical CKs and were poor substrates for a CK oxidase/dehydrogenase, suggesting enhanced biological stability. MeCKs were synthesized by S-adenosyl methionine-dependent methyltransferases (MT1 and MT2) that are present upstream of the fas genes. The best substrate for methylation was isopentenyl diphosphate. MT1 and MT2 catalyzed distinct methylation reactions; only the MT2 product was used by FAS4 to synthesize monomethylated N6-(∆2-isopentenyl)adenine. The MT1 product was dimethylated by MT2 and used as a substrate by FAS4 to produce dimethylated N6-(∆2-isopentenyl)adenine. Chemically synthesized MeCKs were comparable in activity. Our results strongly suggest that MeCKs function as CK mimics and play a role in this plant-pathogen interaction.The balance of phytohormones, such as cytokinins (CKs) and auxins, is finely controlled to maintain proper plant growth and development but is often disturbed following pathogen infection (Robert-Seilaniantz et al., 2007; Pieterse et al., 2012). As a virulence strategy, many phytopathogens synthesize phytohormones that cause aberrant organogenesis and modulate primary carbon metabolism that ultimately aids disease establishment (Jameson, 2000; Robert-Seilaniantz et al., 2007). For several pathogens, CK production is essential for virulence, and they carry genes for CK biosynthesis in a harbored plasmid (Jameson, 2000). Fungal pathogens employ CKs to form green islands with delayed senescence, whereas bacterial pathogens develop gall structures (Sakakibara et al., 2005; Walters et al., 2008; Giron et al., 2013). Rhodococcus fascians is a gram-positive actinomycete that causes symptoms ranging from leaf deformation to differentiated shooty outgrowths known as leafy galls in more than 150 different plant species (Goethals et al., 2001; Stes et al., 2011). In ornamental plants, such infections reduce their value and contribute to economic losses worldwide (Putnam and Miller, 2007). Leafy gall symptoms are reminiscent of CK overproduction and can be partially induced by exogenous application of CKs (Thimann and Sachs, 1966; Eason et al., 1996). Although several CKs have been isolated from R. fascians culture filtrates, a clear correlation with pathogenesis is lacking partially owing to the low concentration of bacterial CKs (Eason et al., 1996). A synergistic action by a mixture of bacterially produced CKs has been proposed, leading to persistent accumulation of CKs locally (Pertry et al., 2009). Nevertheless, to date, no virulence-associated CK analogs have been identified that could contribute to the infection symptoms.Naturally occurring CKs are adenine derivatives with different side chains at the N6 position. Major plant CKs are N6-prenylated adenine derivatives such as N6-(Δ2-isopentenyl)adenine (iP), trans-zeatin (tZ), cis-zeatin (cZ), and dihydrozeatin, collectively known as isoprenoid CKs (Sakakibara, 2006). Among them, iP and tZ are the major CKs in Arabidopsis (Arabidopsis thaliana). iP is synthesized by adenosine phosphate-isopentenyl transferase (IPT) using dimethylallyl diphosphate (DMAPP) and adenosine phosphate as substrates (Sakakibara, 2006). tZ is formed by hydroxylation of the trans-end of the prenyl side chain of the iP nucleotide. CK homeostasis is governed by both biosynthesis and catabolism and has an important regulatory role in plant growth (Sakakibara, 2006; Werner et al., 2006). CK oxidase/dehydrogenase (CKX) is responsible for an irreversible reaction cleaving the unsaturated isoprenoid side chain that results in the formation of adenine and the corresponding aldehyde (Werner et al., 2006). In Arabidopsis, CKs are perceived by a subset of sensory His kinases, ARABIDOPSIS HIS KINASE2 (AHK2) to AHK4, which undergo a His-Asp phosphorelay leading to induction of direct target genes including type-A ARABIDOPSIS RESPONSE REGULATOR (ARR) genes (Kieber and Schaller, 2010). This two-component signaling system has been implicated in mediating basal and pathogen-induced plant immunity (Choi et al., 2010; Argueso et al., 2012). For instance, infection of Arabidopsis plants by R. fascians reportedly activates type-A ARR5 expression with increased expression of AHK3 and AHK4, resulting in mitotic cell divisions that arrest the infected leaves in a meristematic state to establish a nutrient-rich niche (Depuydt et al., 2008, 2009; Pertry et al., 2010; Stes et al., 2011). As the infection progresses, IPT genes are switched off, whereas the expression of all CKX genes are strongly induced in symptomatic tissues (Depuydt et al., 2008).The virulence determinant of R. fascians is located within the fasciation (fas) locus, an operon encoding several genes involved in CK metabolism, indicating that CKs are essential for this plant-pathogen interaction (Stes et al., 2011). fas4 encodes IPT that catalyzes the rate-limiting step of CK biosynthesis and is vital for virulence (Stes et al., 2013). Interestingly, two methyltransferase-like genes are present upstream of the fas gene, whose functions have been unknown. Despite the presence of the fas genes in R. fascians, fewer known CKs have been detected compared with other gall-causing pathogens such as Pantoea agglomerans, Agrobacterium tumefaciens, and Pseudomonas savastanoi (Goethals et al., 2001). Further, the leafy gall phenotype is unique, not invoked by any of the above-mentioned pathogens, implying that the virulence of R. fascians might not be due to typical CKs alone (Goethals et al., 2001). R. fascians has long been hypothesized to produce CK analogs using similar or modified substrates (Goethals et al., 2001; Galis et al., 2005; Stes et al., 2011), but no such molecules have been discovered so far. Here, we report the identification and mode of biosynthesis for methylated cytokinins (MeCKs) as hormone mimics from R. fascians. These compounds are synthesized by two methyltransferases and FAS4. Their CK-like activity and higher in planta stability suggest a role for the methylated analogs as CK mimics that foster efficient pathogenesis.  相似文献   

16.
Brassinosteroid (BR) and gibberellin (GA) are two predominant hormones regulating plant cell elongation. A defect in either of these leads to reduced plant growth and dwarfism. However, their relationship remains unknown in rice (Oryza sativa). Here, we demonstrated that BR regulates cell elongation by modulating GA metabolism in rice. Under physiological conditions, BR promotes GA accumulation by regulating the expression of GA metabolic genes to stimulate cell elongation. BR greatly induces the expression of D18/GA3ox-2, one of the GA biosynthetic genes, leading to increased GA1 levels, the bioactive GA in rice seedlings. Consequently, both d18 and loss-of-function GA-signaling mutants have decreased BR sensitivity. When excessive active BR is applied, the hormone mostly induces GA inactivation through upregulation of the GA inactivation gene GA2ox-3 and also represses BR biosynthesis, resulting in decreased hormone levels and growth inhibition. As a feedback mechanism, GA extensively inhibits BR biosynthesis and the BR response. GA treatment decreases the enlarged leaf angles in plants with enhanced BR biosynthesis or signaling. Our results revealed a previously unknown mechanism underlying BR and GA crosstalk depending on tissues and hormone levels, which greatly advances our understanding of hormone actions in crop plants and appears much different from that in Arabidopsis thaliana.  相似文献   

17.
Accumulating evidence indicates important functions for phosphoenolpyruvate (PEP) carboxylase (PEPC) in inorganic phosphate (Pi)-starved plants. This includes controlling the production of organic acid anions (malate, citrate) that are excreted in copious amounts by proteoid roots of nonmycorrhizal species such as harsh hakea (Hakea prostrata). This, in turn, enhances the bioavailability of mineral-bound Pi by solubilizing Al3+, Fe3+, and Ca2+ phosphates in the rhizosphere. Harsh hakea thrives in the nutrient-impoverished, ancient soils of southwestern Australia. Proteoid roots from Pi-starved harsh hakea were analyzed over 20 d of development to correlate changes in malate and citrate exudation with PEPC activity, posttranslational modifications (inhibitory monoubiquitination versus activatory phosphorylation), and kinetic/allosteric properties. Immature proteoid roots contained an equivalent ratio of monoubiquitinated 110-kD and phosphorylated 107-kD PEPC polypeptides (p110 and p107, respectively). PEPC purification, immunoblotting, and mass spectrometry indicated that p110 and p107 are subunits of a 430-kD heterotetramer and that they both originate from the same plant-type PEPC gene. Incubation with a deubiquitinating enzyme converted the p110:p107 PEPC heterotetramer of immature proteoid roots into a p107 homotetramer while significantly increasing the enzyme’s activity under suboptimal but physiologically relevant assay conditions. Proteoid root maturation was paralleled by PEPC activation (e.g. reduced Km [PEP] coupled with elevated I50 [malate and Asp] values) via in vivo deubiquitination of p110 to p107, and subsequent phosphorylation of the deubiquitinated subunits. This novel mechanism of posttranslational control is hypothesized to contribute to the massive synthesis and excretion of organic acid anions that dominates the carbon metabolism of the mature proteoid roots.Phosphoenolpyruvate (PEP) carboxylase (PEPC; EC 4.1.1.31) is a ubiquitous and tightly regulated cytosolic enzyme of vascular plants that is also widely distributed in green algae and bacteria. PEPC catalyzes the irreversible β-carboxylation of PEP to form oxaloacetate (OAA) and inorganic phosphate (Pi). Vascular plant PEPCs belong to a small multigene family encoding several closely related plant-type PEPCs (PTPCs), along with a distantly related bacterial-type PEPC (BTPC; O’Leary et al., 2011a). PTPC genes encode 105- to 110-kD polypeptides that typically assemble as approximate 400-kD Class-1 PEPC homotetramers. In contrast, BTPC genes encode larger 116- to 118-kD polypeptides owing to a unique intrinsically disordered region that mediates BTPC’s tight interaction with coexpressed PTPC subunits. This association results in the formation of unusual Class-2 PEPC heterooctameric complexes that are largely desensitized to allosteric effectors and that dynamically associate with the surface of mitochondria in vivo (O’Leary et al., 2009, 2011a; Igawa et al., 2010; Park et al., 2012).The critical role of Class-1 PEPC in assimilating atmospheric CO2 during C4 and Crassulacean acid metabolism photosynthesis has been studied extensively. Class-1 PEPCs also fulfill a wide range of crucial nonphotosynthetic functions, particularly the anaplerotic replenishment of tricarboxylic acid cycle intermediates consumed during biosynthesis (O’Leary et al., 2011a). Class-1 PEPCs are subject to a complex set of posttranslational controls including allosteric effectors, covalent modification via phosphorylation or monoubiquitination, and protein-protein interactions (Uhrig et al., 2008; O’Leary et al., 2009, 2011a, 2011b). Allosteric activation by Glc-6-P and inhibition by l-malate are routinely observed, whereas phosphorylation and dephosphorylation are catalyzed by a Ca2+-independent PEPC protein kinase (PPCK) and a protein phosphatase type-2A (PP2A), respectively (O’Leary et al., 2011a). Phosphorylation at a conserved N-terminal seryl residue activates Class-1 PEPCs by decreasing inhibition by malate while increasing activation by Glc-6-P. By contrast, Class-1 PEPC is subject to inhibitory monoubiquitination during castor oil (Ricinus communis) seed (COS) germination, or following depodding of developing COS (Uhrig et al., 2008; O’Leary et al., 2011b). Immunoblots of germinating COS extracts revealed a 1:1 ratio of immunoreactive 110- and 107-kD PTPC polypeptides (p110 and p107, respectively). PEPC purification and mass spectrometry (MS) demonstrated that (1) p110 and p107 are subunits of a 440-kD Class-1 PEPC heterotetramer, (2) both subunits arise from the same PTPC gene (RcPpc3) that also encodes the phosphorylated 410-kD Class-1 PEPC homotetramer of intact developing COS, and (3) p110 is a monoubiquitinated form of p107 (Uhrig et al., 2008). The monoubiquitination site (Lys-628) of COS p110 is conserved in vascular plant PEPCs and is proximal to a PEP-binding/catalytic domain. Incubation with a deubiquitinating enzyme converted the Class-1 PEPC p110:p107 heterotetramer into a p107 homotetramer while exerting significant effects on the enzyme’s kinetic properties (Uhrig et al., 2008). PTPC monoubiquitination rather than phosphorylation is widespread throughout the astor plant and appears to be the predominant posttranslational modification (PTM) of Class-1 PEPC that occurs in unstressed plants (O’Leary et al., 2011b). The distinctive developmental patterns of Class-1 PEPC phosphoactivation versus monoubiquitination-inhibition indicated that these PTMs might be mutually exclusive in the castor plant (O’Leary et al., 2011a, 2011b).Substantial evidence indicates that PEPC plays a pivotal role in plant acclimation to nutritional Pi deficiency (Duff et al., 1989; Vance et al., 2003; O’Leary et al., 2011a; Plaxton and Tran, 2011; Supplemental Fig. S1), a common abiotic stress that frequently limits plant growth in natural ecosystems. The marked induction of Class-1 PEPCs during Pi stress has been linked to the synthesis and excretion of large amounts of organic acid anions by roots of Pi-starved (–Pi) plants (O’Leary et al., 2011a; Uhde-Stone et al., 2003; Vance et al., 2003; Shane et al., 2004a). The excreted organic acids chelate metal cations such as Al3+ and Ca2+ that immobilize Pi in the soil, thus increasing soluble Pi concentrations by up to 1,000-fold (Vance et al., 2003). Harsh hakea (Hakea prostrata) is a perennial nonmycotroph that has evolved a host of traits that allow it to thrive in the nutrient-impoverished, ancient soils of western Australia. A crucial adaptation of harsh hakea is its proteoid roots, which excrete copious quantities of citrate and malate to mediate Pi solubilization and acquisition from the soil’s mineral-bound Pi (Supplemental Figs. S1 and S2; Shane et al., 2003, 2004a, 2004b; Shane and Lambers, 2005). Shane and coworkers (2004a) correlated proteoid root development in –Pi harsh hakea with marked increases in respiration, internal carboxylate concentrations, and rates of carboxylate exudation. Immunoblotting indicated that PEPC abundance remained relatively constant during proteoid root development, except in senescing 3-week-old roots, where it showed a marked decline. The PEPC immunoblots also revealed approximately 110- and 100-kD immunoreactive polypeptides that were of equal intensity in young proteoid roots, whereas mature proteoid roots showed a marked reduction in the p110 (Shane et al., 2004a). The possible contribution of PTMs such as phosphorylation to the in vivo activation of proteoid root PEPCs is currently unclear (e.g. see Uhde-Stone et al., 2003). However, this is feasible since the pronounced induction of PPCK genes coupled with the reversible phosphorylation-activation of a Class-1 PEPC isozyme (AtPPC1) has been conclusively demonstrated in –Pi Arabidopsis (Arabidopsis thaliana) suspension cells and seedlings (Gregory et al., 2009).The goal of the current study was to test the hypothesis that PEPC PTMs contribute to the metabolic adaptations of harsh hakea proteoid roots. We report a novel metabolic control paradigm that involves the in vivo deubiquitination and consequent kinetic activation of a phosphorylated form of a C3 plant Class-1 PEPC.  相似文献   

18.
The C4 photosynthesis carbon-concentrating mechanism in maize (Zea mays) has two CO2 delivery pathways to the bundle sheath (BS; via malate or aspartate), and rates of phosphoglyceric acid reduction, starch synthesis, and phosphoenolpyruvate regeneration also vary between BS and mesophyll (M) cells. The theoretical partitioning of ATP supply between M and BS cells was derived for these metabolic activities from simulated profiles of light penetration across a leaf, with a potential 3-fold difference in the fraction of ATP produced in the BS relative to M (from 0.29 to 0.96). A steady-state metabolic model was tested using varying light quality to differentially stimulate M or BS photosystems. CO2 uptake, ATP production rate (JATP; derived with a low oxygen/chlorophyll fluorescence method), and carbon isotope discrimination were measured on plants under a low light intensity, which is considered to affect C4 operating efficiency. The light quality treatments did not change the empirical ATP cost of gross CO2 assimilation (JATP/GA). Using the metabolic model, measured JATP/GA was compared with the predicted ATP demand as metabolic functions were varied between M and BS. Transamination and the two decarboxylase systems (NADP-malic enzyme and phosphoenolpyruvate carboxykinase) were critical for matching ATP and reduced NADP demand in BS and M when light capture was varied under contrasting light qualities.Interest in the C4 pathway has been increased by the potential for enhancing crop productivity and maintaining yield stability in the face of global warming and population pressure (Friso et al., 2010; Zhu et al., 2010; Covshoff and Hibberd, 2012). Maize (Zea mays), a C4 plant of the NADP-malic enzyme (ME) subtype, is a leading grain production cereal (www.fao.org). C4 photosynthesis is a shared activity between mesophyll (M; abbreviations are listed in BS) cells, coupled to allow the operation of a biochemical carbon-concentrating mechanism (CCM). The CCM effectively minimizes photorespiration by increasing the CO2 concentration in the bundle sheath (CBS), where Rubisco is exclusively expressed. Since BS and M are connected by plasmodesmata, some CO2 retrodiffuses. The refixation of that escaping CO2 by the CCM increases the activity of the CCM and the total ATP demand (ATPBS + ATPM) for gross CO2 assimilation (GA; [ATPBS + ATPM]/GA), from a theoretical minimum of five ATPs (Furbank et al., 1990). Leakiness (Φ), the amount of CO2 retrodiffusing relative to phosphoenolpyruvate (PEP) carboxylation rate, is therefore a proxy for the coordination between the CCM and assimilatory activity (Henderson et al., 1992; Tazoe et al., 2008; Kromdijk et al., 2010; Ubierna et al., 2011; Bellasio and Griffiths, 2013).

Table I.

Variables and acronyms described in the text
AbbreviationDefinitionUnit
ANet assimilationμmol m−2 s−1
ABAbsorbed light
AB BS/MPartitioning of absorbed lightDimensionless
ATPBSATP demand in BSμmol m−2 s−1
ATPMATP demand in Mμmol m−2 s−1
BSBundle sheath
CBSCO2 concentration in BSμmol mol−1
CCMCarbon-concentrating mechanism
CEFCyclic electron flow
DHAPDihydroxyacetone phosphate
ETRElectron transport rateμmol m−2 s−1
GAGross assimilation (A + RLIGHT)μmol m−2 s−1
gBSBundle sheath conductance to CO2, calculated by fitting JMOD to JATPmol m2 s−1
IRGAInfrared gas analyzer
JATPTotal ATP production rateμmol m−2 s−1
JATPBSATP production rate in BSμmol m−2 s−1
JATPMATP production rate in Mμmol m−2 s−1
JMODModeled ATP production rateμmol m−2 s−1
LEFLinear electron flow
LCPLight compensation point
MMesophyll
MALMalate
MDHMalate dehydrogenase
MDHBSMalate dehydrogenase reaction rate in BSμmol m−2 s−1
MDHMMalate dehydrogenase reaction rate in Mμmol m−2 s−1
MEMalic enzyme
MEMalic enzyme reaction rateμmol m−2 s−1
NADPHBSNADPH demand in BSμmol m−2 s−1
NADPHTOTTotal NADPH demandμmol m−2 s−1
OAAOxaloacetic acid
PARPhotosynthetically active radiationμE m−2 s−1
PEPPhosphoenolpyruvate
PEPCKPhosphoenolpyruvate carboxykinase
PEPCKPEPCK reaction rateμmol m−2 s−1
PGA3-Phosphoglyceric acid
PPDKPyruvate phosphate dikinase
PPDKPPDK reaction rateμmol m−2 s−1
PRPGA reduction
PRBSPR rate in BSμmol m−2 s−1
PRMPR rate in Mμmol m−2 s−1
RBSRespiration in the light in BSμmol m−2 s−1
RLIGHTRespiration in the lightμmol m−2 s−1
RPPReductive pentose phosphate
RuBPRibulose-1,5-bisphosphate
RuPRibulose-5-phosphate
SSStarch synthesis
SSBSStarch synthesis rate in BSμmol m−2 s−1
SSMStarch synthesis rate in Mμmol m−2 s−1
SSTOTTotal starch synthesis rateμmol m−2 s−1
TTransamination rateμmol m−2 s−1
VCRubisco carboxylation rateμmol m−2 s−1
VORubisco oxygenation rateμmol m−2 s−1
VPPEP carboxylation rateμmol m−2 s−1
Y(II)Yield of PSII
Δ13C isotopic discrimination
δ13C13C isotopic composition relative to Pee Dee Belemnite
ΦLeakinessDimensionless
Open in a separate windowRecently, the maize C4 subgroup has been shown to be complicated by the presence of two BS decarboxylation enzyme systems (NADP-ME and phosphoenolpyruvate carboxykinase [PEPCK]), presumably both acting as CO2 delivery pathways (via malate [MAL] and Asp, respectively; Furumoto et al., 1999, 2000; Wingler et al., 1999; Eprintsev et al., 2011; Furbank, 2011; Pick et al., 2011). There is also an extensive overlap between BS and M functions, since both cell types can synthesize starch (Spilatro and Preiss, 1987; Kanai and Edwards, 1999) and reduce phosphoglyceric acid (PGA; Majeran and van Wijk, 2009; see the overall scheme in Fig. 1). Additionally, energetic partitioning can also vary between cell types, since the total ATP produced (JATP) per CO2 fixed in GA (JATP/GA) may be produced in BS (mainly through cyclic electron flow [CEF] around PSI) or in M (mainly through linear electron flow [LEF]), depending on the light locally available in BS or M (Kramer and Evans, 2011; Yin and Struik, 2012). Furthermore, although all NADPH is produced in M, the only compartment operating linear electron transport and oxidizing water, some NADPH is exported to BS through MAL diffusion, to meet the reducing power demand therein (NADPHBS). To capture the complex C4 physiology, several models of C4 photosynthesis have been developed (Berry and Farquhar, 1978; Laisk and Edwards, 2000, 2009; von Caemmerer, 2000). The earlier approaches were developed into the von Caemmerer (2000) C4 model. In particular, the associated light-limited equations (referred to subsequently as the “C4 model”) are used to estimate the parameters needed to resolve the isotopic discrimination (Δ) model, widely employed to study Φ under low-light conditions (for review, see Ubierna et al., 2011). The C4 model partitions JATP into two fractions: (1) the ATP consumed by PEP carboxylase, and (2) the ATP consumed by the C3 activity (glyoxylate recycling, PGA reduction [PR], and ribulose 1,5-bisphosphate [RuBP] regeneration). These activities are located in M, BS, or both compartments (see the overall scheme in Fig. 1). However, the C4 model simplifies the spatial compartmentalization between BS and M, and in this paper, we now develop the energetic implications of the differential contribution of M and BS to C4 photosynthesis under different light regimes.Open in a separate windowFigure 1.Metabolic model of C4 assimilation, rates of reaction, and net fluxes between BS and M. The overall scheme reports the reactions of the CCM (Furbank, 2011), Rubisco carboxylation, the reactions of the RPP pathway, the synthesis of starch, respiration, and glyoxylate recycling reactions. The tables, with the corresponding enzyme names, show the actual reaction rates, expressed relative to GA (5.13 μmol m−2 s−1), per unit of substrate transformed. Rates were estimated by parameterizing the model equations (PAR = 125 μE m−2 s−1 (A = 3.96 μmol m−2 s−1; RLIGHT = 1.17 μmol m−2 s−1; JATP = 28.6 μmol m−2 s−1), the output of the C4 model (VC = 5.35 μmol m−2 s−1; VP = 5.89 μmol m−2 s−1; VO = 0.44 μmol m−2 s−1), and the output of the Δ model (Φ = 0.23) under three characteristic ratios of ATP partitionings. These were numbered 1, 2, and 3. Condition 1 corresponds to the lowest ATP available in BS (ATP partitioning similar to that under blue light; Fig. 4B), condition 2 corresponds to an intermediate ATP availability in BS (ATP partitioning equal to that under red light; Fig. 4B), and condition 3 corresponds to the highest ATP available in BS (ATP partitioning equal to that under green light; Fig. 4B). The inset shows net metabolite fluxes between M and BS in multiples of GA. The ATP demand in BS (ATPBS) and M (ATPM), the total NADPH demand (NADPHTOT), and the NADPHBS were also calculated in the same three relevant conditions. PYR, Pyruvic acid.Because of these anatomical, metabolic, and energetic complexities, C4 metabolism is highly sensitive to limiting light intensity (Bellasio and Griffiths, 2013) and, potentially, light quality (Evans et al., 2007). Light quality has a greater influence on C4 photosynthesis than on C3. Leaf pigments preferentially absorb the blue and red region of the spectra, and some wavelengths penetrate deeper into leaves. It was shown in C3 leaves that exposure to different wavelengths results in characteristic light penetration profiles, which, translated into different gradients in PSII yield, rates of ATP production, and assimilation (A) within the leaf (Terashima et al., 2009). In C4 leaves, because of the concentric anatomy, light reaches M cells before the deeper BS (Evans et al., 2007) and could alter the balance between light harvesting and energetic partitioning between BS and M.In this paper, we model the likely profiles of light penetration for specific wavelengths associated with red, green, and blue light within a maize M and BS leaf cross section and calculate the impact on potential ATP production for each cell type. We calculate the proportion of absorbed light (AB) for each wavelength, expressed as AB BS/M, the fraction of photons absorbed in BS relative to the photons absorbed in M, from which we derive JATPBS/JATPM, the fraction of ATP produced in BS relative to the ATP produced in M. Second, we developed a steady-state metabolic model (Fig. 1; von Caemmerer 2000), to capture the spatial separation between BS and M and partitions the ATP demand between BS and M cells in terms of PR, starch synthesis (SS), and PEP regeneration, so as to meet the ATP availability in each cell type (Evans et al., 2007). Third, photosynthetic characteristics (leaf-level ATP production rate, CO2 assimilation, stomatal conductance, and Φ derived from online carbon isotope discrimination [Δ]) were measured under red, green, and blue light, and red, green, and blue light in combination (RGB), using a decreasing photon flux density (from 500 to 50 μE m−2 s−1) to investigate the importance of metabolic plasticity under limiting light intensities.

Table II.

Steady-state equations for the metabolic model of C4 assimilationProcesses described by Equations 4 to 10 can be calculated directly from the measured data for A, RLIGHT, and the output of the von Caemmerer C4 model (VO, VP, and VC), while Equations 11 to 21 require prior allocation of SS, PR, and PEPCK. For simplicity, enzyme names in italics represent the enzyme reaction rate. For stoichiometric consistency, reaction rates are calculated as rates of substrate transformation.
ProcessSymbolReaction RateEquationLocalizationNotes
Gross assimilationGA(4)GA and RLIGHT rates are expressed per CO2.
RuP phosphorylation(5)BSRuP phosphorylation supplies Rubisco carboxylating activity (VC) together with oxygenating activity (VO).
Total PRPRTOT(6)BS and MThis equation calculates the total rate of PR on the basis of the PGA produced by Rubisco carboxylation (2VC), Rubisco oxygenation (VO), and glyoxylate recycling (0.5VO) and considers the PGA consumed by respiration; 1/3 is the stoichiometric conversion between respiration (expressed per CO2) and PR (expressed per triose).
Total NADPH demandNADPHTOT(7)BS and MPR consumes one NADPH per PGA; the total rate of PR is PRTOT (see note to Eq. 6); in glyoxylate regeneration (per glyoxylate), 0.5 NADH is produced by Gly decarboxylase, 0.5 NADH is consumed by hydroxypyruvate reductase, and one ferredoxin (equivalent to 0.5 NADPH) is consumed by Gln synthetase; in total, 0.5 NADPH is consumed per glyoxylate (equivalent to VO rate; Supplemental Table S1; Yoshimura et al., 2004).
DHAP entering RPP(8)BSThe DHAP entering the RPP pathway corresponds to the total PR rate minus the DHAP used for starch synthesis, which in this work is expressed per triose.
Total SSSSTOT(9)BS and MIn this model, assimilation is entirely converted to starch; this assumption does not influence energetics, as starch synthesis has the same ATP demand as phloem-loaded Suc; in Equation 9, 1/3 converts the stoichiometry of A (expressed per CO2) to the stoichiometry of SS (expressed per triose).
Total PEP regeneration(10)BS and MPEP regeneration rate equals PEP consumption rate VP at steady state; PEP can be regenerated either by PPDK (mainly in M but active also in BS) or by PEPCK in BS; in this study, PPDK activity was assumed to be zero in BS.
Total ATP demandATPBS + ATPM(11)BS and MEquation 11 calculates the total ATP demand as the sum of ATP demand for PR (one ATP per PGA, corresponding to PR), RuBP regeneration (one ATP per RuP, corresponding to VC + VO), glyoxylate recycling (one ATP per glyoxylate, corresponding to VO), starch synthesis (0.5 ATP per triose, corresponding to SS), and PEP regeneration (one ATP per PEPCK catalytic event or two ATP per PPDK catalytic event); compared with the original formulation of the C4 model, Equation 11 separates the ATP demand for PEPCK and PPDK, includes the ATP demand for SS, and considers the PGA utilized by respiration, which does not need to be reduced (see Eq. 6).
ATP demand in BSATPBS(12)BSThe ATP demand in BS is brought about by PR (at the rate of PRBS), RuBP regeneration (at the rate of VC + VO), glyoxylate recycling (at the rate of VO), starch synthesis (0.5 ATP per triose), and PEPCK activity (one ATP per OAA; see note to Eq. 11).
ATP demand in MATPM(13)MThe ATP demand in M is brought about by PR (at the rate of PRM), SS, and PPDK (two ATPs per pyruvic acid; see note to Eq. 11).
NADPH demand in BSNADPHBS(14)BSThe NADPH demand in BS is brought about by PR (one NADPH per PGA) and glyoxylate recycling, which consumes 0.5 NADPH per glyoxylate (corresponding to VO; see Supplemental Table S2).
NADPH supply to BSMDHM(15)BSAll NADPH available in BS is produced in M and exported through the MAL shuttle because we have assumed that no linear electron transport (i.e. water oxidation) occurred in BS; for this reason, the NADPH supply to BS corresponds to the NADPH consumed to reduce OAA to MAL in M, the process responsible for NADPH export, and not to the rate of MAL decarboxylation in BS, which depends on T, PEPCK, and MDHBS (Eq. 19).
MDH activity in MMDHM(16)MMDH activity supplies the NADPH demand in BS; Equation 16 was derived from Equations 14 and 15.
TransaminationT(17)BS and MEquation 17 expresses that, at steady state, all OAA is either transaminated or reduced; since T bypasses the MDHM reaction, which is the reaction responsible for NADPH export to BS (see note to Eq. 15), T has the function of balancing NADPH supply and demand, which becomes apparent when Equations 15 and 17 are combined.
MDHMDHBST − PEPCK(18)BSMDH is assumed to operate a fast conversion at equilibrium; therefore, it is passively regulated by the substrate availability: the OAA that is not used by PEPCK is reduced to MAL by MDH; MDH may use NADH, since no NADPH-dependent reduction of OAA has been observed in maize (Kanai and Edwards, 1999) and it is likely mitochondrial (Rathnam, 1978; Chapman and Hatch, 1981); the NADH regeneration may be carried out by chloroplastic ME, which is reported to react both with NADP and NAD (Chapman and Hatch, 1981); however, the process may be more complicated (Eprintsev et al., 2011, and refs. therein); note that in this study, we assumed that cells are decompartmentalized while PEPCK rate was manipulated to increase between zero and a maximum rate in response to ATP availability (see “Minimum and Maximum BS Allocation” for details).
MEMEMDHM + MDHBS19BSEquation 19 expresses that the rate of MAL oxidation by ME corresponds to the rate of MAL produced by MDH activity in M plus the rate of MAL produced by MDH activity in BS.
PPDKPPDKVP − PEPCK20MThe PEP regenerated by PEPCK in BS diffuses to M and reduces the requirement of PEP regenerated by PPDK in M.
PR in MPRMPRTOT − PRBS21MPR is a shared process between BS and M.
Open in a separate windowFor instance, AB BS/M and JATPBS/JATPM were both lower under the blue light (wavelength 460 nm), which is rapidly extinguished within the M leaf profile, than under white light, confirming that light quality perturbs C4 energetics. In spite of this shift, when maize plants were exposed to different light qualities, there was no change in Φ, indicating that, at steady state, the coordination between CCM activity and Rubisco assimilation was retained (Ubierna et al., 2011; Sun et al., 2012). The modeled metabolic plasticity projected a window for ATP demand partitioning (ATPBS/ATPM), which matched the values for JATPBS/JATPM supply estimated under red, green, and blue light. We show that the plasticity of C4 metabolism, and in particular the possibility of shifting between MAL and Asp as a primary carboxylase product, was of pivotal importance in allowing the plasticity of ATP and NADPH demand. In conclusion, our study explains the extensive overlap between BS and M functions and the requirement for at least two decarboxylase systems in NADP-ME subtype plants such as maize, providing an explanation for empirical observations on the diversity of decarboxylase activities and PEP regeneration pathways (Rathnam, 1978; Chapman and Hatch, 1981; Wingler et al., 1999; Eprintsev et al., 2011; Furbank, 2011; Pick et al., 2011).  相似文献   

19.
Stresses decouple nitrate assimilation and photosynthesis through stress-initiated nitrate allocation to roots (SINAR), which is mediated by the nitrate transporters NRT1.8 and NRT1.5 and functions to promote stress tolerance. However, how SINAR communicates with the environment remains unknown. Here, we present biochemical and genetic evidence demonstrating that in Arabidopsis thaliana, ethylene (ET) and jasmonic acid (JA) affect the crosstalk between SINAR and the environment. Electrophoretic mobility shift assays and chromatin immunoprecipitation assays showed that ethylene response factors (ERFs), including OCTADECANOID-RESPONSIVE ARABIDOPSIS AP2/ERF59, bind to the GCC boxes in the NRT1.8 promoter region, while ETHYLENE INSENSITIVE3 (EIN3) binds to the EIN3 binding site motifs in the NRT1.5 promoter. Genetic assays showed that cadmium and sodium stresses initiated ET/JA signaling, which converged at EIN3/EIN3-Like1 (EIL1) to modulate ERF expression and hence to upregulate NRT1.8. By contrast, ET and JA signaling mediated the downregulation of NRT1.5 via EIN3/EIL1 and other, unknown component(s). SINAR enhanced stress tolerance and decreased plant growth under nonstressed conditions through the ET/JA-NRT1.5/NRT1.8 signaling module. Interestingly, when nitrate reductase was impaired, SINAR failed to affect either stress tolerance or plant growth. These data suggest that SINAR responds to environmental conditions through the ET/JA-NRT signaling module, which further modulates stress tolerance and plant growth in a nitrate reductase-dependent manner.  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