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.
Prolonged dry periods, and increasingly the generation of smoke and dust in partially-deforested regions, can influence the chemistry of rainfall and throughfall in moist tropical forests. We investigated rainfall and throughfall chemistry in a palm-rich open tropical rainforest in the southwestern Brazilian Amazon state of Rondônia, where precipitation averages 2300 mm year?1 with a marked seasonal pattern, and where the fragmentation of remaining forest is severe. Covering the transition from dry to wet season (TDWS) and the wet season (WS) of 2004–2005, we sampled 42 rainfall events on event basis as well as 35 events on a within-event basis, and measured concentrations of DOC, Na+, K+, Ca2+, Mg2+, NH 4 + , Cl?, SO 4 2? , NO 3 ? and pH in rainfall and throughfall. We found strong evidence of both seasonal and within-event solute rainfall concentration dynamics. Seasonal volume-weighted mean (VWMS) concentrations in rainfall of DOC, K+, Ca2+, Mg2+, NH 4 + , SO 4 2? and NO 3 ? were significantly higher in the TDWS than the WS, while VWMS concentrations in throughfall were significantly higher for all solutes except DOC. Patterns were generally similar within rain events, with solute concentrations declining sharply during the first few millimeters of rainfall. Rainfall and throughfall chemistry dynamics appeared to be strongly influenced by forest and pasture burning and a regional atmosphere rich in aerosols at the end of the dry season. These seasonal and within-event patterns of rainfall and throughfall chemistry were stronger than those recorded in central Amazônia, where the dry season is less pronounced and where regional deforestation is less severe. Fragmentation and fire in Rondônia now appear to be altering the patterns in which solutes are delivered to remaining moist tropical forests.  相似文献   

2.
Calmodulin binds to IQ motifs in the α1 subunit of CaV1.1 and CaV1.2, but the affinities of calmodulin for the motif and for Ca2+ are higher when bound to CaV1.2 IQ. The CaV1.1 IQ and CaV1.2 IQ sequences differ by four amino acids. We determined the structure of calmodulin bound to CaV1.1 IQ and compared it with that of calmodulin bound to CaV1.2 IQ. Four methionines in Ca2+-calmodulin form a hydrophobic binding pocket for the peptide, but only one of the four nonconserved amino acids (His-1532 of CaV1.1 and Tyr-1675 of CaV1.2) contacts this calmodulin pocket. However, Tyr-1675 in CaV1.2 contributes only modestly to the higher affinity of this peptide for calmodulin; the other three amino acids in CaV1.2 contribute significantly to the difference in the Ca2+ affinity of the bound calmodulin despite having no direct contact with calmodulin. Those residues appear to allow an interaction with calmodulin with one lobe Ca2+-bound and one lobe Ca2+-free. Our data also provide evidence for lobe-lobe interactions in calmodulin bound to CaV1.2.The complexity of eukaryotic Ca2+ signaling arises from the ability of cells to respond differently to Ca2+ signals that vary in amplitude, duration, and location. A variety of mechanisms decode these signals to drive the appropriate physiological responses. The Ca2+ sensor for many of these physiological responses is the Ca2+-binding protein calmodulin (CaM).2 The primary sequence of CaM is tightly conserved in all eukaryotes, yet it binds and regulates a broad set of target proteins in response to Ca2+ binding. CaM has two domains that bind Ca2+ as follows: an amino-terminal domain (N-lobe) and a carboxyl-terminal domain (C-lobe) joined via a flexible α-helix. Each lobe of CaM binds two Ca2+ ions, and binding within each lobe is highly cooperative. The two lobes of CaM, however, have distinct Ca2+ binding properties; the C-lobe has higher Ca2+ affinity because of a slower rate of dissociation, whereas the N-lobe has weaker Ca2+ affinity and faster kinetics (1). CaM can also bind to some target proteins in both the presence and absence of Ca2+, and the preassociation of CaM in low Ca2+ modulates the apparent Ca2+ affinity of both the amino-terminal and carboxyl-terminal lobes. Differences in the Ca2+ binding properties of the lobes and in the interaction sites of the amino- and carboxyl-terminal lobes enable CaM to decode local versus global Ca2+ signals (2).Even though CaM is highly conserved, CaM target (or recognition) sites are quite heterogeneous. The ability of CaM to bind to very different targets is at least partially due to its flexibility, which allows it to assume different conformations when bound to different targets. CaM also binds to various targets in distinct Ca2+ saturation states as follows: Ca2+-free (3), Ca2+ bound to only one of the two lobes, or fully Ca2+-bound (47). In addition, CaM may bind with both lobes bound to a target (5, 6) or with only a single lobe engaged (8). If a target site can bind multiple conformers of CaM, CaM may undergo several transitions that depend on Ca2+ concentration, thereby tuning the functional response. Identification of stable intermediate states of CaM bound to individual targets will help to elucidate the steps involved in this fine-tuned control.Both CaV1.1 and CaV1.2 belong to the L-type family of voltage-dependent Ca2+ channels, which bind apoCaM and Ca2+-CaM at carboxyl-terminal recognition sites in their α1 subunits (914). Ca2+ binding to CaM, bound to CaV1.2 produces Ca2+-dependent facilitation (CDF) (14). Whether CaV1.1 undergoes CDF is not known. However, both CaV1.2 and CaV1.1 undergo Ca2+- and CaM-dependent inactivation (CDI) (14, 15). CaV1.1 CDI is slower and more sensitive to buffering by 1,2-bis(o-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid than CaV1.2 CDI (15). Ca2+ buffers are thought to influence CDI and/or CDF in voltage-dependent Ca2+ channels by competing with CaM for Ca2+ (16).The conformation of the carboxyl terminus of the α1 subunit is critical for channel function and has been proposed to regulate the gating machinery of the channel (17, 18). Several interactions of this region include intramolecular contacts with the pore inactivation machinery and intermolecular contacts with CaM kinase II and ryanodine receptors (17, 1922). Ca2+ regulation of CaV1.2 may involve several motifs within this highly conserved region, including an EF hand motif and three contiguous CaM-binding sequences (10, 12). ApoCaM and Ca2+-CaM-binding sites appear to overlap at the site designated as the “IQ motif” (9, 12, 13), which are critical for channel function at the molecular and cellular level (14, 23).Differences in the rate at which 1,2-bis(o-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid affects CDI of CaV1.1 and CaV1.2 could reflect differences in their interactions with CaM. In this study we describe the differences in CaM interactions with the IQ motifs of the CaV1.1 and the CaV1.2 channels in terms of crystal structure, CaM affinity, and Ca2+ binding to CaM. We find the structures of Ca2+-CaM-IQ complexes are similar except for a single amino acid change in the peptide that contributes to its affinity for CaM. We also find that the other three amino acids that differ in CaV1.2 and CaV1.1 contribute to the ability of CaV1.2 to bind a partially Ca2+-saturated form of CaM.  相似文献   

3.
Kinetics of the reduction of the hemes in cytochrome c oxidase in the presence of high concentration of ruthenium(III)hexaammine chloride was examined using a stopped-flow spectrophotometer. Upon mixing of the oxidized enzyme with dithionite and Ru(NH3) 6 3+ , three well-resolved phases were observed: heme a reduction reaching completion within a few milliseconds is followed by two slow phases of heme a 3 reduction. The difference spectrum of heme a 3 reduction in the visible region is characterized by a maximum at ~612 nm, rather than at 603 nm as was believed earlier. It is shown that in the case of bovine heart cytochrome c oxidase containing a special cation-binding site in which reversible binding of calcium ion occurs, heme a 3 reduction is slowed down by low concentrations of Ca2+. The effect is absent in the case of the bacterial cytochrome oxidase in which the cation-binding site contains a tightly bound Ca2+ ion. The data corroborate the inhibition of the cytochrome oxidase enzymatic activity by Ca2+ ions discovered earlier and indicate that the cation affects intramolecular electron transfer.  相似文献   

4.
The electromagnetic field (EMF) is newly considered as an exogenous environmental stimulus that is closely related to ion transportation on the cellular membrane, maintaining the internal ionic homeostasis. Cation transports of Ca2+ and other metal ions, Cd2+, Zn2+, and Mn2+were studied in terms of the external Ca2+ stress, [Ca2+]ext, and exposure to the physical EMF. A specific yeast strain K667 was used for controlling CAX5 (cation/H+ exchanger) expression. Culture samples were exposed to 60 Hz, 0.1 mT sinusoidal or square magnetics waves, and intracellular cations of each sample were measured and analyzed. AtCAX5 transformant yeast grew normally under the metallic stress. However, the growth of the control group was significantly inhibited under the same cation concentration; 60 Hz and 0.1 mT magnetic field enhanced intracellular cation concentrations significantly as exposure time increased both in the AtCAX5 transformed yeast and in the control group. However, the AtCAX5-transformed yeast showed higher concentration of the intracellular cations than the control group under the same exposure EMF. AtCAX5-transformed yeasts displayed an increment in [Ca2+]int, [K+]int, [Na+]int, and [Zn2+]int concentration under the presence of both sinusoidal and square-waved EMF stresses compared to the control group, which shows that AtCAX5 expressed in the vacuole play an important role in maintaining the homeostasis of intracellular cations. These findings could be utilized in the cultivation of the crops which were resistant to excessive exogenous ions or in the production of biomass containing a large proportion of ions for nutritional food or in the bioremediation process in metal-polluted environments.  相似文献   

5.
Ninety-one lakes distributed along the Tatra Mountains (most of lakes > 1 ha and 65% of lakes > 0.01 ha) were sampled and analysed for ionic and nutrient composition in September 2004 (15 years after reduction in acid deposition). Eighty-one lakes were in alpine zone and ten lakes in Norway spruce forest. The results were compared to similar lake surveys from 1994 (the beginning of water recovery from acidification) and 1984 (maximum acidification). Atmospheric deposition of SO 4 2? and inorganic N decreased 57% and 35%, respectively, in this region from the late 1980s to 2000. Lake water concentrations of SO 4 2? and NO 3 ? have decreased both by ~50% on average (to 23 and 19 μmol L?1, respectively, in 2004) since 1984. While the decrease in SO 4 2? concentrations was stable throughout 1984–2004, most of the NO 3 ? decrease occurred from 1994 to 2004. The declines in SO 4 2? and NO 3 ? concentrations depended on catchment coverage with vegetation, being most rapid for SO 4 2? in forest lakes and for NO 3 ? in rocky lakes. Concentrations of the sum of base cations (dominated by Ca2+) significantly decreased between 1984 and 2004, with the highest change in rocky lakes. Most of this decline occurred between 1994 and 2004. Acid neutralising capacity (ANC) did not change in the 1984–1994 period, but increased on average by 29 μmol L?1 between 1994 and 2004, with the highest change in rocky lakes. Over the last decade, the proportion of lakes with ANC > 150 μmol L?1 increased from 15% to 21% and that of ANC < 20 μmol L?1 decreased from 37% to 20%. The highest decline in H+ and Al concentrations occurred in the most acid lakes. On a regional basis, no significant change was observed for total phosphorus, total organic nitrogen, and dissolved organic carbon (DOC) in the 1994–2004 period. However, these parameters increased in forest lakes, which exhibited an increasing trend in DOC concentrations, inversely related (P < 0.001) to their decreasing ionic strength (30% on average in 1994–2004).  相似文献   

6.
The first cytogenetic comparisons of five snapper species from Thailand were presented here. Renal cell samples were taken from blacktail snapper (Lutjanus fulvus), five lined snapper (L. quinquelineatus), dory snapper (L. fulviflamma), brownstripe red snapper (L. vitta), and mangrove red snapper (L. argentimaculatus). The mitotic chromosome preparation was prepared directly from kidney cells. Conventional staining and Ag-NOR banding techniques were applied to stain the chromosomes. The results exhibited that all five snapper species have the diploid chromosome numbers of 2n = 48 and the fundamental numbers (NF) of 48. The presences of large, medium, and small telocentric chromosomes were 22-24-2, 24-20-4, 36-10-2, 28-16-4 and 36-10-2, respectively. The Ag- NORs banding technique provides the pair of nucleolar organizer regions (NORs) at subcentromeric region of the long arm of the respective telocentric chromosome pairs 9, 1, 3, 4 and 9. Their karyotype formulas is as follows: L. fulvus (2n = 48): L 22 t + M 24 t + S 2 t , L. quinquelineatus (2n = 48): L 24 t + M 20 t + S 4 t , L. fulviflamma (2n = 48): Lt36 + Mt10 + St2, L. vitta (2n = 48): L 28 t + M 16 t + S 4 t , and L. argentimaculatus (2n = 48): L 36 t + M 10 t + S 2 t .  相似文献   

7.
Reactive oxygen species (ROS) and calcium (Ca2+), two crucial intracellular signaling molecules, have been reported to play important roles in chlorophyll biosynthesis. In this study, we aimed to investigate whether disturbance of chlorophyll synthesis affects chloroplast ROS and Ca2+ homeostases. Chlorophyll biosynthesis was inhibited at the Mg branch by virus-induced gene silencing (VIGS) of CHLI gene encoding the Mg chelatase CHLI subunit in pea (Pisum sativum). Subsequently, ROS and intracellular free Ca2+ concentration ([Ca2+]i) in these chlorophyll-deficient pea plants were evaluated by histochemical and fluorescent staining assays. The results showed that the superoxide anion and hydrogen peroxide were predominantly generated in chloroplasts of the yellow leaves of pea VIGS-CHLI plants. The expression of genes encoding chloroplast antioxidant enzymes (CuZn-superoxide dismutase, ascorbate peroxidase, glutathione reductase, phospholipid glutathione peroxidase, peroxiredoxin and thioredoxins) were also decreased in the leaves of VIGS-CHLI plants compared with the control plants. Additionally, the [Ca2+]i were significantly reduced in the yellow leaves of VIGS-CHLI plants compared with the green leaves of VIGS-GFP control plants. The expression of genes encoding Ca2+ signaling related proteins (thylakoid Ca2+ transporter, calmodulins and calcineurin B-like protein) was down-regulated in yellow VIGS-CHLI leaves. These results indicate that inhibition of chlorophyll biosynthesis at the Mg branch by silencing CHLI affects chloroplast ROS homeostasis and Ca2+ signaling and down-regulates the expression of ROS scavenging genes and Ca2+ signaling related genes.  相似文献   

8.
Interactions between metal ions and amino acids are common both in solution and in the gas phase. The effect of metal ions and water on the structure of l-histidine is examined. The effect of metal ions (Li+, Na+, K+, Mg2+, Ca2+, Ni2+, Cu2+ and Zn2+) and water on structures of His·M(H2O)m, m = 0.1 complexes have been determined theoretically employing density functional theories using extended basis sets. Of the five stable complexes investigated the relative stability of the gas-phase complexes computed with DFT methods (with one exception of K+ systems) suggest metallic complexes of the neutral l-histidine to be the most stable species. The calculations of monohydrated systems show that even one water molecule has a profound effect on the relative stability of individual complexes. Proton dissociation enthalpies and Gibbs energies of l-histidine in the presence of the metal cations Li+, Na+, K+, Mg2+, Ca2+, Ni2+, Cu2+ and Zn2+ were also computed. Its gas-phase acidity considerably increases upon chelation. Of the Lewis acids investigated, the strongest affinity to l-histidine is exhibited by the Cu2+ cation. The computed Gibbs energies ΔG are negative, span a rather broad energy interval (from ?130 to ?1,300 kJ/mol), and upon hydration are appreciably lowered.  相似文献   

9.
Calcium ion (Ca2+) is essential secondary messenger in plant signaling networks. In this study, the effect of Ca2+ on oxidative damage caused by a high irradiance (HI) was investigated in the leaves of two cultivars of tall fescue (Arid3 and Houndog5). Pretreatment of the tall fescue leaves with a CaCl2 solution significantly increased Ca2+ content and intrinsic HI tolerance due to a decreased ion leakage and content of malondialdehyde, hydrogen peroxide, and superoxide radicals. Moreover, the activities of superoxide dismutase, catalase, ascorbate peroxidase, and glutathione reductase increased in both the cultivars in the presence of Ca2+ under the HI stress. In contrast, treatments with a Ca2+ chelator ethylene glycol-bis(2-aminoethylether)-N,N,N′,N′-tetraacetic acid (EGTA) or a plasma membrane Ca2+ channel blocker LaCl3 reversed these effects. On the other hand, a pronounced increase in nitric oxide synthase-like activity and NO release by exogenous Ca2+ treatment was observed in the tolerant Arid3 plants after exposure to the HI, whereas only a small increase was observed in more sensitive Houndog5. Moreover, the inhibition of NO production by 2-(4-carboxy-2-phenyl)-4,4,5,5-tetramethylimidazoline-1-oxyl-3-oxide or Nω-nitro-L-arginine blocked the protective effect of exogenous Ca2+, whereas the inhibition of Ca2+ by EGTA or LaCl3 had no influence on the protective effect of NO. The results indicate that NO might be involved in the Ca2+-induced activities of antioxidant enzymes further protecting against HI-induced oxidative damage. This protective mechanism was found to be more efficient in Arid3 than in Houndog5.  相似文献   

10.
11.
  1. The affinity of ATP-supported Ca2+ accumulation for both Ca2+ and ATP was determined from initial rate studies employing isolated rat liver mitochondria. TheK m values for “free” Ca2+ and ATP were calculated to be of the order of 2 μM and 100 μM, respectively. TheK m for ATP decreased as the Ca2+ concentration was increased.
  2. The curve relating initial rates of Ca2+ accumulation to Ca2+ concentration was singmoidal in shape; values obtained for the Hill coefficient were in the range 1.5–1.9.
  3. Concomitant with the ATP-stimulated accumulation of Ca2+, ATP translocation was itselt increased in the presence of Ca2+. This stimulation took place independently of Ca2+ accumulation.
  4. Decreasing the pH of the incubation medium decreased the rate of Ca2+ accumulation. This inhibition was competitive in that the affinity of mitochondrial for Ca2+ could be altered. The maximal rate of accumulation did not change with change in pH.
  5. The permeant anions inorganic phosphate and acetate stimulated the accumulation of Ca2+ in a non-competitive manner. Both theV max and Km varied when either of the anions were present.
  6. The data are discussed in relation to the role that mitochondria play in controlling the cellular ionic environment.
  相似文献   

12.
The antifungal protein PAF from Penicillium chrysogenum exhibits growth-inhibitory activity against a broad range of filamentous fungi. Evidence from this study suggests that disruption of Ca2+ signaling/homeostasis plays an important role in the mechanistic basis of PAF as a growth inhibitor. Supplementation of the growth medium with high Ca2+ concentrations counteracted PAF toxicity toward PAF-sensitive molds. By using a transgenic Neurospora crassa strain expressing codon-optimized aequorin, PAF was found to cause a significant increase in the resting level of cytosolic free Ca2+ ([Ca2+]c). The Ca2+ signatures in response to stimulation by mechanical perturbation or hypo-osmotic shock were significantly changed in the presence of PAF. BAPTA [bis-(aminophenoxy)-ethane-N,N,N′,N′-tetraacetic acid], a Ca2+ selective chelator, ameliorated the PAF toxicity in growth inhibition assays and counteracted PAF induced perturbation of Ca2+ homeostasis. These results indicate that extracellular Ca2+ was the major source of these PAF-induced effects. The L-type Ca2+ channel blocker diltiazem disrupted Ca2+ homeostasis in a similar manner to PAF. Diltiazem in combination with PAF acted additively in enhancing growth inhibition and accentuating the change in Ca2+ signatures in response to external stimuli. Notably, both PAF and diltiazem increased the [Ca2+]c resting level. However, experiments with an aequorin-expressing Δcch-1 deletion strain of N. crassa indicated that the L-type Ca2+ channel CCH-1 was not responsible for the observed PAF-induced elevation of the [Ca2+]c resting level. This study is the first demonstration of the perturbation of fungal Ca2+ homeostasis by an antifungal protein from a filamentous ascomycete and provides important new insights into the mode of action of PAF.The secreted antifungal protein PAF from Penicillium chrysogenum is a small-molecular-mass (6.2 kDa), cationic, and cysteine-rich peptide that inhibits the growth of numerous filamentous fungi (1416, 21). It belongs to a family of antifungal peptides which show—despite considerable amino acid homology—significant differences in species specificity and modes of action (reviewed in reference 27). Importantly, the primary structures of these antifungals show no similarity to those from higher eukaryotes, e.g., plants, insects, or mammals (see reference 28 for a detailed review on parallels with and differences between PAF and antimicrobial proteins from higher eukaryotes and their mechanisms of action). Apart from the Aspergillus giganteus-derived antifungal protein AFP (19, 45; reviewed in reference 29), PAF is one of the best-studied peptides of this protein family. We have shown that PAF causes rapid hyperpolarization of the plasma membrane at hyphal tips, increased K+ efflux, induction of oxidative stress, and apoptotic cell death (21, 25) and that PAF is internalized by hyphae of PAF-sensitive fungi (33). Furthermore, we have shown that PAF interferes with at least two signaling cascades, the protein kinase C/mitogen-activated protein (MAP) kinase and the cyclic AMP (cAMP)/protein kinase A pathways, which play a role in mediating PAF toxicity (5). However, it still has to be elucidated in more detail how these PAF-dependent effects are linked.Recent evidence indicated that the ionic strength of the growth medium interferes with the antifungal activity of PAF (21). It is possible that Ca2+ ions may play a major role in influencing protein toxicity in an analogous way to the Ca2+-dependent mode of action of antifungal plant defensins. Supplementation of the test medium with low concentrations of CaCl2 (1 to 5 mM) reversed the antifungal activity of plant defensins (34, 4344). The defensins RsAFP2 from the seeds of Raphanus sativus and DmAMP1 from the seeds of Dahlia merckii induced K+ efflux and Ca2+ uptake in Neurospora crassa and caused alkalinization of the growth medium (46). Another seed defensin, MsDef1 from Medicago sativa, was reported to cause Ca2+ influx and the inhibition of mammalian L-type Ca2+ channels, similar to the Ustilago maydis killer toxin KP4 (13, 40).Ca2+ is a universal intracellular second messenger in eukaryotic cells (4, 36). In fungi, there is evidence for Ca2+ signaling regulating numerous processes, including spore germination, tip growth, hyphal branching, sporulation, infection structure differentiation, and circadian clocks, as well as responses to osmotic stress, heat shock, mechanical stimuli, oxidative stress, and electrical fields (7, 11, 17, 20, 22, 3031, 3839, 41, 49). Ca2+ signaling typically involves transient increases in intracellular Ca2+ concentrations originating from the extracellular medium and/or mobilization from internal stores (4, 23, 3637). Little was known about Ca2+ dynamics in filamentous fungi until a new method based on the Ca2+-sensitive photoprotein aequorin was adapted, allowing routine and easy measurement of intracellular calcium dynamics in these organisms (3, 30).The aim of our study was to investigate the effect of PAF on the level of cytosolic free Ca2+ ([Ca2+]c) in the PAF-sensitive target organism N. crassa to characterize its effects on stimulus-specific Ca2+ signatures and to define the Ca2+ sources responsible for the perturbation of Ca2+ homeostasis. By obtaining [Ca2+]c measurements in living cells expressing the bioluminescent Ca2+ reporter aequorin, we provide novel insights into the mode of action of this biotechnologically promising antifungal protein.  相似文献   

13.
Fluorescence resonance energy transfer (FRET)-sensitized emission imaging of Arabidopsis thaliana roots expressing the yellow cameleon 3.60 calcium (Ca2+) reporter showed that the concentration of calcium in the cytosol ([Ca2+]cyt) increased upon aluminum ion (Al3+) treatment in root cells from the transition zone within seconds. The Al3+-induced [Ca2+]cyt transients were biphasic and were modified by Ca2+ channel blockers and by an antagonist of neuronal glutamate receptors, 2-amino-5-phosphonopentanoate (AP-5), and by the anion channel blocker, 5-nitro-2-(3′-phenylpropyl-amino) benzoate (NPPB). The [Ca2+]cyt transients were not uniquely associated with Al3+ toxicity mechanisms since lanthanum (La3+) and gadolinium (Gd3+) also elicited [Ca2+]cyt transients that were similar to those induced by Al3+. Here a testable model that describes a possible mechanism and sequence of events that lead to the Al3+-induced [Ca2+]cyt transients and inhibition of root growth is proposed. This model can be applied to study also the signal-response coupling of the trivalent ions La3+ and Gd3+.Key words: aluminum toxicity, Al3+ transport, Ca2+ signaling, fluorescence resonance energy transfer (FRET), yellow cameleonAluminum (Al) is a naturally occurring component of soil particles and is the third most abundant element in the earth''s crust.1 In acidic soils, Al dissolves in the soil solution and different ionic Al species form.2,3 The most toxic Al species in acidic soils is ionic Al, Al3+.4 Al3+ toxicity stems from its interference with a plethora of cellular processes that control plant growth and development.3,57The interactions between calcium (Ca2+) and Al3+ are well documented in the literature. One of the toxic effects of Al3+ on plant growth and development has been ascribed to the disruption of Ca2+ homeostasis by Al3+.8,9 The fact that Al3+ inhibits Ca2+ uptake by roots,10 blocks voltage-regulated Ca2+ channels,11,12 and affects the concentration of Ca2+ in the cytosol ([Ca2+]cyt)1318 support this view. Ca2+ alleviates Al3+ toxicity1922 perhaps by inhibiting Al3+ accumulation in the roots and cells.23,24Rincón-Zachary et al.18 using fluorescence resonance energy transfer (FRET)-sensitized emission to image Arabidopsis thaliana roots expressing the yellow cameleon 3.60 Ca2+ reporter demonstrated increases in the concentration of free Ca2+ in the cytosol ([Ca2+]cyt) within seconds of Al3+ application. Al3+ induced distinct [Ca2+]cyt signatures in cells from the different developmental root regions-meristem, elongation and maturation zones. The [Ca2+]cyt signature in the transition zone, which is the most Al-sensitive root region,25 was biphasic and was modified by treatments that chelate external Ca2+ (EGTA), block Ca2+ entry through the plasma membrane (verapamil), by an antagonist of neuronal glutamate receptors, 2-amino-5-phosphonopentanoate (AP-5), and by the anion channel blocker, 5-nitro-2-(3′-phenylpropyl-amino) benzoate (NPPB). All of these agents affected the first peak of the Al3+-induced [Ca2+]cyt signature by reducing its magnitude or abolishing it. These results support the notion that Al3+ interacts with different types of plasma membrane Ca2+ channels, causing them to open. Al3+-induced [Ca2+]cyt transients were also observed in the Arabidopsis Al-resistant and Al-sensitive mutants alr104 and als3, respectively. In addition, the trivalent ions lanthanum (La3+) and gadolinium (Gd3+) evoked [Ca2+]cyt signatures in the transition zone of the wild-type Arabidopsis and of the alr104 and als3 roots similar to those elicited by Al3+. Hence the authors concluded that the observed [Ca2+]cyt transients were not uniquely associated with Al3+ toxicity mechanisms. Al3+, La3+ and Gd3+ appear to elicit the same Ca2+ signaling pathway.I would like to propose a testable model that describes the possible sequence of events during Ca2+ signaling triggered by trivalent ions using Al3+ as a prototype (Fig. 1). (1) Al3+ causes Ca2+ channels in cells of the root transition zone to open allowing Ca2+ influx into the cells. (2) [Ca2+]cyt rises producing the first peak of the biphasic [Ca2+]cyt signature. (3) Increased [Ca2+]cyt activates internal Ca2+ channels located in membranes of internal Ca2+ stores such as the vacuole, ER, mitochondria or plastids producing the second peak of the [Ca2+]cyt signature. Ca2+-induced Ca2+ release from internal stores has been described in plant cells.26 (4) Al3+ may permeate plasma membrane Ca2+ and non-selective cation channels and interact with internal Ca2+ channels allowing Ca2+ to be released into the cytosol, contributing to the rise in [Ca2+]cyt. In this context, supporting data come from unpublished results (Leblanc J and Rincón-Zachary M) that show Al3+ transport across plasma membrane (PM) vesicles isolated from 5 mm wheat (Triticum aestivum) root tips by aqueous two-phase partitioning27 (Fig. 2). In this experiment isolated PM vesicles were loaded with the fluorescent histochemical aluminum indicator morin (2′, 3′, 4′, 5, 7-pentahydroxyflavone) for 30 min at room temperature and then centrifuged at 100,000 xg for 15 min at 4°C and the pellet was washed twice to remove excess morin. The PM vesicles (25 µg protein mL−1) were then incubated in a 2 mL buffer (250 mM sucrose, 50 mM K2SO4, 1 mM DTT, 5 mM MES-Tris [pH 7.0]) containing different concentrations of Al3+ for 10 min at room temperature. Al3+uptake by the PM vesicles was monitored by fluorometry (excitation at 420 nm; emission at 475 nm). The results show that PM vesicles isolated from the Al-sensitive wheat cultivar Scout 66 root tips are more permeable to Al3+ than those isolated from the Al-tolerant cultivar Atlas 66 (Fig. 2A). In this experiment, the relationship between the rate of Al3+ uptake and the Al3+ concentration in the solution was linear for both Scout 66 (Y = 0.114X + 0.741, R2 = 0.99) and Atlas 66 (Y = 0.108X + 0.193, R2 = 0.98) PM vesicles. In addition, Leblanc28 showed that compounds known to block Ca2+ channels inhibited Al3+ uptake by plasma membrane vesicles (Fig. 2B; Leblanc J and Rincón-Zachary M, unpublished data). La3+, verapamil and nifedipine were very effective in inhibiting Al3+ uptake by plasma membrane vesicles: 5 µM La3+ and 1 mM nifedipine caused 67% and 73% inhibition, respectively, and 1 mM verapamil completely abolished the Al3+ uptake by the vesicles. Thus, it is feasible that Al3+ permeates non-selective cation channels or/and Ca2+ channels. (5) Lastly, the overall [Ca2+]cyt elevation could set off mechanisms that inhibit root growth (e.g., callose synthesis and its deposition in the cell wall, disruption of the cytoskeleton organization, formation of reactive oxygen species, etc.). Testing these hypotheses is underway.Open in a separate windowFigure 1A model that describes a possible mechanism and sequence of events that lead to the [Ca2+]cyt transients and inhibition of root growth. (1) Al3+ interacts with Ca2+ channels in the plasma membrane of root cells in the root transition zone. The Ca2+channels open and external Ca2+ enters the cytosol. (2) [Ca2+]cyt rises producing the first peak of the biphasic [Ca2+]cyt signature. (3) Increased [Ca2+]cyt activates internal Ca2+ channels located in membranes of internal Ca2+ stores (e.g., tonoplast, ER, mitochondria or plastids) producing the second peak of the [Ca2+]cyt signature. (4) Al3+ permeates the PM through Ca2+- and non-selective cation channels. (5) Al3+ opens internal Ca2+ channels in the tonoplast, ER, mitochondria or plastids and as a result more Ca2+ is released into the cytosol. (6) The overall [Ca2+]cyt elevation stimulates mechanisms that inhibit root growth.Open in a separate windowFigure 2Al3+ uptake by PM vesicles isolated from 5 mm root tips of both the Al-sensitive cultivar Scout 66 and the Al-tolerant cultivar Atlas 66. (A) Rate of Al3+ uptake by PM vesicles incubated in increasing concentrations of Al3+. The PM vesicles from the Al sensitive cultivar Scout 66 were more permeable to Al3+ than those of the Al-tolerant cultivar Atlas 66. The values are means ± SD. Rates of Al3+ uptake are expressed in Fluorescence Intensity Units (FIU) mg−1 protein min−1. (B) Effect of Ca2+ channel blockers on the rate of Al3+ uptake by PM vesicles s percent of the control. All Ca2+ channel blockers tested inhibited the rate Al3+ uptake by the PM vesicles in both cultivars. The accumulation of Al3+ in the PM vesicles was monitored by measuring the fluorescence emitted by the Al-morin complex as described in the text. Both experiments were repeated three times in triplicate (n = 9). The PM vesicles were pooled from multiple independent membrane isolations in order to obtain enough membrane protein for the assays.  相似文献   

14.
Fluorescence resonance energy transfer-sensitized emission of the yellow cameleon 3.60 was used to study the dynamics of cytoplasmic calcium ([Ca2+]cyt) in different zones of living Arabidopsis (Arabidopsis thaliana) roots. Transient elevations of [Ca2+]cyt were observed in response to glutamic acid (Glu), ATP, and aluminum (Al3+). Each chemical induced a [Ca2+]cyt signature that differed among the three treatments in regard to the onset, duration, and shape of the response. Glu and ATP triggered patterns of [Ca2+]cyt increases that were similar among the different root zones, whereas Al3+ evoked [Ca2+]cyt transients that had monophasic and biphasic shapes, most notably in the root transition zone. The Al3+-induced [Ca2+]cyt increases generally started in the maturation zone and propagated toward the cap, while the earliest [Ca2+]cyt response after Glu or ATP treatment occurred in an area that encompassed the meristem and elongation zone. The biphasic [Ca2+]cyt signature resulting from Al3+ treatment originated mostly from cortical cells located at 300 to 500 μ m from the root tip, which could be triggered in part through ligand-gated Glu receptors. Lanthanum and gadolinium, cations commonly used as Ca2+ channel blockers, elicited [Ca2+]cyt responses similar to those induced by Al3+. The trivalent ion-induced [Ca2+]cyt signatures in roots of an Al3+-resistant and an Al3+-sensitive mutant were similar to those of wild-type plants, indicating that the early [Ca2+]cyt changes we report here may not be tightly linked to Al3+ toxicity but rather to a general response to trivalent cations.The role of calcium ions (Ca2+) as a ubiquitous cellular messenger in animal and plant cells is well established (Berridge et al., 2000; Sanders et al., 2002; Ng and McAinsh, 2003). Cellular signal transduction pathways are elicited as a result of fluctuations of free Ca2+ in the cytoplasm ([Ca2+]cyt) in response to external and intracellular signals. These changes in [Ca2+]cyt influence numerous cellular processes, including vesicle trafficking, cell metabolism, cell proliferation and elongation, stomatal opening and closure, seed and pollen grain germination, fertilization, ion transport, and cytoskeletal organization (Hepler, 2005). [Ca2+]cyt fluctuations occur because cells have a Ca2+ signaling “toolkit” (Berridge et al., 2000) composed of on/off switches and a multitude of Ca2+-binding proteins. The on switches depend on membrane-localized Ca2+ channels that control the entry of Ca2+ into the cytosol (Piñeros and Tester, 1995, 1997; Thion et al., 1998; Kiegle et al., 2000a; White et al., 2000; Demidchik et al., 2002; Miedema et al., 2008). On the other hand, the off switches consist of a family of Ca2+-ATPases and Ca2+/H+ exchangers in the plasma membrane or endomembrane that remove Ca2+ from the cytosol, bringing the [Ca2+]cyt down to the initial resting level (Lee et al., 2007; Li et al., 2008).The numerous cellular processes regulated by Ca2+ have led investigators to ask how specificity in Ca2+ signaling is maintained. It has been proposed that specificity in Ca2+ signaling is achieved because a particular stimulus elicits a distinct Ca2+ signature, which is defined by the timing, magnitude, and frequency of [Ca2+]cyt changes. For instance, tip-growing plant cells such as root hairs and pollen tubes exhibit oscillatory elevations in [Ca2+]cyt that partly mirror the oscillatory nature of growth in these cell types (Cárdenas et al., 2008; Monshausen et al., 2008). Another example is nuclear Ca2+ spiking in root hairs of legumes exposed to NOD factors (Oldroyd and Downie, 2006; Peiter et al., 2007). Recently, it was shown that mechanical forces applied to an Arabidopsis (Arabidopsis thaliana) root can trigger a stimulus-specific [Ca2+]cyt response (Monshausen et al., 2009). Translating the Ca2+ signature into a defined cellular response is governed by a number of Ca2+-binding proteins such as calreticulin that act as [Ca2+]cyt buffers, which shape both the amplitude and duration of the Ca2+ signal or Ca2+ sensors such as calmodulin that impact other downstream cellular effectors (Berridge et al., 2000; White and Broadley, 2003; Hepler, 2005).A deeper understanding of Ca2+ signaling mechanisms in plants has been driven in large part by our ability to monitor dynamic changes in [Ca2+]cyt in the cell. Such measurements have been conducted using Ca2+-sensitive fluorescent indicator dyes (e.g. Indo and Fura), the luminescent protein aequorin (Knight et al., 1991, 1996; Legué et al., 1997; Wymer et al., 1997; Cárdenas et al., 2008), and more recently the yellow cameleon (YC) Ca2+ sensor, a chimeric protein that relies on fluorescence resonance energy transfer (FRET) as an indicator of [Ca2+]cyt changes in the cell (Allen et al., 1999; Miwa et al., 2006; Qi et al., 2006; Tang et al., 2007; Haruta et al., 2008). The YC reporter is composed of cyan fluorescent protein (CFP), the C terminus of calmodulin (CaM), a Gly-Gly linker, the CaM-binding domain of myosin light chain kinase (M13), and a yellow fluorescent protein (YFP; Miyawaki et al., 1997, 1999). The increased interaction between M13 and CaM upon binding of Ca2+ to CaM triggers a conformational change in the protein that brings the CFP and YFP in close proximity, resulting in enhanced FRET efficiency between the two fluorophores (Miyawaki, 2003). Thus, changes in FRET efficiency between CFP and YFP in the cameleon reporter are correlated with changes in [Ca2+]cyt.Since it was first introduced, improved versions of the cameleon reporter have been selected to more accurately report [Ca2+]cyt levels in the cell. For instance, the YC3.60 version was selected because of its resistance to cytoplasmic acidification and its higher dynamic range compared with the earlier cameleons. The higher dynamic range of YC3.60 is due to the use of a circularly permutated YFP called Venus (cpVenus) that is capable of absorbing a greater amount of energy from CFP (Nagai et al., 2004). Recently, the utility of YC3.60 for monitoring [Ca2+]cyt was demonstrated in Arabidopsis roots and pollen tubes using ratiometric imaging approaches (Monshausen et al., 2007, 2008, 2009; Haruta et al., 2008; Iwano et al., 2009). Here, we further evaluated YC3.60 as a [Ca2+]cyt sensor in plants using confocal microscopy and FRET-sensitized emission imaging. Unlike the direct ratiometric measurement of cpVenus and CFP reported in previous studies using YC3.60-expressing plants (Monshausen et al., 2008, 2009), the sensitized FRET approach we describe here involves the use of donor-only (CFP) and acceptor-only (YFP) controls, allowing us to correct for bleed-through and background signals from the FRET specimen (van Rheenen et al., 2004; Feige et al., 2005).For this study, we focused on monitoring [Ca2+]cyt changes in Arabidopsis seedling roots after aluminum (Al3+) exposure. Although Ca2+ signaling has long been implicated in mediating Al3+ responses in plants (Rengel and Zhang, 2003), the [Ca2+]cyt changes evoked by Al3+ reported in the literature have been inconsistent, and as such, the significance of these [Ca2+]cyt responses to mechanisms of Al3+ toxicity are not very clear. For instance, some studies reported that Al3+ caused a decrease in [Ca2+]cyt in plants (Jones et al., 1998b; Kawano et al., 2004), and others demonstrated elevated [Ca2+]cyt upon Al3+ treatment (Nichol and Oliveira, 1995; Lindberg and Strid, 1997; Jones et al., 1998a; Zhang and Rengel, 1999; Ma et al., 2002; Bhuja et al., 2004).Here, we report that Arabidopsis roots expressing the YC3.60 reporter exhibited transient elevations in [Ca2+]cyt within seconds of Al3+ exposure. The general pattern of [Ca2+]cyt changes observed after Al3+ treatment were distinct from those elicited by ATP or Glu, reinforcing the concept of specificity in [Ca2+]cyt signaling. We also observed root zone-dependent variations in the [Ca2+]cyt signatures evoked by Al3+ in regard to the shape, duration, and timing of the [Ca2+]cyt response. Other trivalent ions such as lanthanum (La3+) and gadolinium (Ga3+), which have been widely used as Ca2+ channel blockers (Monshausen et al., 2009), also induced a rapid rise in [Ca2+]cyt in root cells that were similar to those elicited by Al3+. Al3+, La3+, and Gd3+ elicited similar [Ca2+]cyt signatures in the Al3+-tolerant mutant alr104 (Larsen et al., 1998) and the Al3+-sensitive mutant als3-1 (Larsen et al., 2005), indicating that the early [Ca2+]cyt increases we report here may not be tightly linked to mechanisms of Al3+ toxicity but rather to a general trivalent cation response. Our study further shows that FRET-sensitized emission imaging of Arabidopsis roots expressing YC3.60 provides a robust method for documenting [Ca2+]cyt signatures in different root developmental zones that should be useful for future studies on Ca2+ signaling mechanisms in plants.  相似文献   

15.
16.
Transport of electrons in spinach photosystem II (PSII) whose oxygen-evolving complex (OEC) contains heterogeneous metal clusters 2Mn2Fe and 3Mn1Fe was studied by measuring the fluorescence induction kinetics (FIK). PSII(2Mn,2Fe) and PSII(3Mn,1Fe) preparations were produced using Cadepleted PSII membranes (PSII(–Ca)). It was found that FIK in PSII(2Mn,2Fe) membranes is similar in form to FIK in PSII(–Ca) samples, but the fluorescence yield is lower in PSII(2Mn,2Fe). The results demonstrate that, just as in PSII(–Ca) preparations, there is electron transfer from the metal cluster in the OEC to the primary plastoquinone electron acceptor QA. They also show that partial substitution of Mn cations with Fe has no effect on the electron transport on the acceptor side of PSII. Thus, these data demonstrate the possibility of water oxidation either by the heterogeneous metal cluster or just by the manganese dimer. We established that FIK in PSII(3Mn,1Fe) preparations are similar in form to FIK in PSII(2Mn,2Fe) membranes but PSII(3Mn,1Fe) is characterized by a slightly higher maximal fluorescence yield, Fmax. The electron transfer rate in PSII(3Mn,1Fe) preparations significantly (by a factor of two) increases in the presence of Ca2+, whereas Ca2+ has hardly any effect on the electron transport in PSII(2Mn,2Fe) membranes. In Mndepleted PSII membranes, FIK reaches its maximum (the so-called peak K), after which the fluorescence yield starts to decrease as the result of two factors: the oxidation of reduced primary plastoquinone Q A ? and the absence of electron influx from the donor side of PSII. The replacement of Mn cations by Fe in PSII(?Mn) preparations leads to fluorescence saturation and disappearance of the K peak. This is possibly due to the deceleration of the charge recombination process that takes place between reduced primary electron acceptor Q A ? and oxidized tyrosine Y Z +. which is an electron carrier between the OEC and the primary electron donor P680.  相似文献   

17.
Brevibacteria able to decrease phosphate concentration in the medium are of interest for the study of the role of bacteria in the phosphorus cycle and for development of biotechnology of phosphate removal from waste. Brevibacterium casei, Brevibacterium linens, and Brevibacterium epidermidis grown in media with initial phosphorus concentrations of 1–11 mM were shown to decrease its concentration by 90%. The composition of the incubation medium required for B. casei to carry out this process was established. This process occurs in the absence of glucose but requires the presence of Mg2+, NH 4 + , and α-ketoglutarate. The latter two components may be replaced by amino acids metabolized to NH 4 + and α-ketoglutarate: histidine, arginine, glutamine, proline, or glutamic acid. No formation of insoluble phosphate salts was observed when the media were incubated under the same conditions with heat-inactivated cells or without cells at pH 7–8.5.  相似文献   

18.
Senescence and wilting of the leaves of pea (Pisum sativum L.) of normal (AfAf) and aphyllous (afaf) genotypes were accompanied by DNA degradation. In young (12th–9th) subapical leaves of AfAf plants, total DNA was high-polymeric; in the 6th leaf, DNA degradation was appreciable; and in the 4th and 3rd leaves, hydrolysis of DNA was pronounced. Similar degradation of DNA was also observed in senescing leaves of aphyllous plants, but there it started later than in the plants of normal type. The extent of DNA degradation was closely related to the elevation of total nuclease activity in pea leaves associated with the age. The leaves of plants of different genotypes distinctly differed in the activity of acid and alkaline nucleases. Senescence of the leaves was accompanied by the induction of Ca2+-and Mg2+-dependent nucleases with mol wts of 42, 37, 34, 26, and 21 kD. In different stages of leaf senescence, different sets of nucleases were detected.  相似文献   

19.
20.
Under the upland forests growing on the Dnieper loamy soils at 150–160 m above sea level, sandy aquifers are found at ~120 and ~140 m above sea level. The general mineralization of groundwaters (GWs) and the concentration of SO 4 -2 increase upon approaching the foundation of fluvial glacial sediments. The concentration of C supplied to the local GWs is 10–12 mg L–1 per year–1. The concentration of C(CO 3 -2 ; HCO 3 -1 ; NO 3 -1 ; NO 2 -1 ) in the DSW is consistent with the C balance of upper biogeocenoses and the rate of groundwater migration. The concentrations of N(nh 4 +1 ) are unrelated to the filtration of biogeocenosis moisture. They are identical both under sodic grassy glades and highly productive tree stands. The main part of N-containing ions in situ is synthesized and destroyed by anaerobic microorganisms living in the GW.  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