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.
NH4+ inhibition kinetics for CH4 oxidation were examined at near-atmospheric CH4 concentrations in three upland forest soils. Whether NH4+-independent salt effects could be neutralized by adding nonammoniacal salts to control samples in lieu of deionized water was also investigated. Because the levels of exchangeable endogenous NH4+ were very low in the three soils, desorption of endogenous NH4+ was not a significant factor in this study. The Km(app) values for water-treated controls were 9.8, 22, and 57 nM for temperate pine, temperate hardwood, and birch taiga soils, respectively. At CH4 concentrations of ≤15 μl liter−1, oxidation followed first-order kinetics in the fine-textured taiga soil, whereas the coarse-textured temperate soils exhibited Michaelis-Menten kinetics. Compared to water controls, the Km(app) values in the temperate soils increased in the presence of NH4+ salts, whereas the Vmax(app) values decreased substantially, indicating that there was a mixture of competitive and noncompetitive inhibition mechanisms for whole NH4+ salts. Compared to the corresponding K+ salt controls, the Km(app) values for NH4+ salts increased substantially, whereas the Vmax(app) values remained virtually unchanged, indicating that NH4+ acted by competitive inhibition. Nonammoniacal salts caused inhibition to increase with increasing CH4 concentrations in all three soils. In the birch taiga soil, this trend occurred with both NH4+ and K+ salts, and the slope of the increase was not affected by the addition of NH4+. Hence, the increase in inhibition resulted from an NH4+-independent mechanism. These results show that NH4+ inhibition of atmospheric CH4 oxidation resulted from enzymatic substrate competition and that additional inhibition that was not competitive resulted from a general salt effect that was independent of NH4+.Atmospheric CH4 contributes substantially to the greenhouse effect, and the concentration of atmospheric CH4 has increased dramatically in the past century because of human activity associated with agriculture, land use changes, and industry (34, 35). Bacterial oxidation of atmospheric CH4 in well-drained soils is an important regulator of atmospheric CH4 concentration, yet the organisms responsible remain unidentified and the physiology of the process is poorly understood (9, 35, 36). Although soil CH4 consumption is inhibited by a wide variety of anthropogenic disturbances, such as agriculture, N deposition, and forestry (12, 17, 22, 23, 32, 43, 44), predictable inhibition patterns have failed to emerge, which has made it difficult to predict the effects of disturbance on soil CH4 flux in various ecosystems. The most commonly reported disturbance effect is that of NH4+ fertilizers, which can suppress soil CH4 consumption by up to 70% (1, 8, 10, 17, 22, 32, 33, 37, 38, 43). In the field, inhibition may occur immediately following fertilization, may be delayed for months to years, or may never occur despite years of chronic fertilization (9, 17). This variety of responses may stem at least in part from the distribution of physiologically diverse methane oxidizer populations across sites (17, 18, 20).Of the various NH4+ inhibition patterns, immediate inhibition is the best documented. As in field studies, however, physiological laboratory studies have produced variable results, suggesting that there may be multiple inhibition mechanisms (15, 17, 2628, 36, 39). Physicochemical similarities between CH4 and NH3 may permit these two compounds to compete for enzyme active sites so that fortuitous NH3 oxidation competitively inhibits CH4 oxidation (38). Although this mechanism has been demonstrated to occur in pure cultures of methanotrophic bacteria (6) and in a CH4-producing agricultural soil (15), it has not been demonstrated to occur in well-drained, nonagricultural mineral soils, which comprise the dominant terrestrial sink for atmospheric CH4 (14, 38, 45), nor has it been demonstrated to occur at near-atmospheric CH4 concentrations. In many cases, the kinetics of immediate NH4+ inhibition in soil cannot be reconciled easily with substrate competition (15, 16, 2628, 39). An alternative mechanism has been proposed, whereby the toxicity of NO2 or NH2OH produced by fortuitous NH4+ oxidation suppresses methanotrophic activity (26, 27, 39). Hence, multiple inhibition mechanisms may be involved, and these mechanisms may vary with the physiology of different CH4 oxidizer populations (17).Two physiologically distinct communities of CH4 oxidizers apparently exist in soil. One group, generally associated with atmospheric CH4 consumption, exhibits an extremely high affinity for CH4. Representatives of this group have yet to be cultivated or otherwise identified (9). The second group exhibits a much lower affinity for CH4 and is generally associated with common methanotrophs, such as those that have been studied in pure culture for many years (2, 9). In upland mineral soils, only high-affinity activity is usually detectable without artificial enrichment with high CH4 concentrations in the laboratory. However, the only prior study in which kinetic constants for NH4+ inhibition of soil CH4 oxidation were reported was conducted in a periodically moist, organic matter-rich agricultural soil with demonstrable methanogenesis (15, 16). Such a soil potentially harbors a rich community of CH4 oxidizers representing a continuum from low-affinity organisms to high-affinity organisms. Although this important investigation demonstrated that NH4+ inhibits CH4 oxidation via enzymatic substrate competition in an agricultural humisol, it is unclear to what extent its results apply to well-drained mineral soils lacking endogenous CH4 sources. Physiological studies of soil CH4 oxidation typically derive kinetic constants from oxidation rates at CH4 concentrations ranging from atmospheric levels (∼1.7 μl liter−1) to ≫Km for high-affinity CH4 oxidizers. Even in soil in which only high-affinity organisms are active, the CH4-oxidizing enzyme(s) could respond differently to NH4+ at high CH4 concentrations than at near-atmospheric concentrations (15, 39). Thus, to study NH4+ inhibition of high-affinity CH4 oxidizers per se, it would be preferable to examine inhibition kinetics at near-atmospheric CH4 concentrations in a soil with no apparent endogenous CH4 source.A common shortcoming of NH4+ inhibition studies, regardless of the organisms involved, has been a lack of attention to nonammoniacal salt effects despite numerous reports of substantial inhibition by such salts (1, 10, 15, 17, 24). King and Schnell (28) examined the effects of several Cl and SO42− salts and concluded that nonammoniacal salts indirectly inhibit CH4 oxidation by desorbing endogenous NH4+ from cation exchange sites in the soil, which then directly inhibit CH4 oxidation. Many N-limited soils, however, have extremely low concentrations of exchangeable NH4+, yet are substantially inhibited by nonammoniacal salts (17), suggesting that these salts have NH4+-independent effects on atmospheric CH4 oxidizers. Additional mechanisms may alter inhibition kinetics, thus hindering the diagnosis of NH4+-specific inhibition.With the limitations described above in mind, we used a simple steady-state kinetics approach to assess the mechanism of NH4+ inhibition of CH4 oxidation at near-atmospheric concentrations (1.8 to 15 μl liter−1) in three well-drained, N-limited forest soils that lack known endogenous CH4 sources. In addition, we examined the effects of nonammoniacal salts in parallel samples to judge the utility of these salts as experimental controls for neutralizing NH4+-independent salt effects.  相似文献   

2.
Methanotrophic and nitrifying bacteria are both able to oxidize CH4 as well as NH4+. To date it is not possible to estimate the relative contribution of methanotrophs to nitrification and that of nitrifiers to CH4 oxidation and thus to assess their roles in N and C cycling in soils and sediments. This study presents new options for discrimination between the activities of methanotrophs and nitrifiers, based on the competitive inhibitor CH3F and on recovery after inhibition with C2H2. By using rice plant soil as a model system, it was possible to selectively inactivate methanotrophs in soil slurries at a CH4/CH3F/NH4+ molar ratio of 0.1:1:18. This ratio of CH3F to NH4+ did not affect ammonia oxidation, but methane oxidation was inhibited completely. By using the same model system, it could be shown that after 24 h of exposure to C2H2 (1,000 parts per million volume), methanotrophs recovered within 24 h while nitrifiers stayed inactive for at least 3 days. This gave an “assay window” of 48 h when only methanotrophs were active. Applying both assays to model microcosms planted with rice plants demonstrated a major contribution of methanotrophs to nitrification in the rhizosphere, while the contribution of nitrifiers to CH4 oxidation was insignificant.  相似文献   

3.
The growth rate of Chromohalobacter salexigens DSM 3043 can be stimulated in media containing 0.3 M NaCl by a 0.7 M concentration of other salts of Na+, K+, Rb+, or NH4+, Cl, Br, NO3, or SO42− ions. To our knowledge, growth rate stimulation by a general high ion concentration has not been reported for any organism previously.  相似文献   

4.
The effect of short-term ammonia starvation on Nitrosospira briensis was investigated. The ammonia-oxidizing activity was determined in a concentrated cell suspension with a NOx biosensor. The apparent half-saturation constant [Km(app)] value of the NH3 oxidation of N. briensis was 3 μM NH3 for cultures grown both in continuous and batch cultures as determined by a NOx biosensor. Cells grown on the wall of the vessel had a lower Km(app) value of 1.8 μM NH3. Nonstarving cultures of N. briensis showed potential ammonia-oxidizing activities of between 200 to 250 μM N h−1, and this activity decreased only slowly during starvation up to 10 days. Within 10 min after the addition of fresh NH4+, 100% activity was regained. Parallel with activity measurements, amoA mRNA and 16S rRNA were investigated. No changes were observed in the 16S rRNA, but a relative decrease of amoA mRNA was observed during the starvation period. During resuscitation, an increase in amoA mRNA expression was detected simultaneously. The patterns of the soluble protein fraction of a 2-week-starved culture of N. briensis showed only small differences in comparison to a nonstarved control. From these results we conclude that N. briensis cells remain in a state allowing fast recovery of ammonia-oxidizing activity after addition of NH4+ to a starved culture. Maintaining cells in this kind of active state could be the survival strategy of ammonia-oxidizing bacteria in nature under fluctuating NH4+ availability.  相似文献   

5.
Techniques of compartmental (efflux) and kinetic influx analyses with the radiotracer 13NH4+ were used to examine the adaptation to hypoxia (15, 35, and 50% O2 saturation) of root N uptake and metabolism in 3-week-old hydroponically grown rice (Oryza sativa L., cv IR72) seedlings. A time-dependence study of NH4+ influx into rice roots after onset of hypoxia (15% O2) revealed an initial increase in the first 1 to 2.5 h after treatment imposition, followed by a decline to less than 50% of influx in control plants by 4 d. Efflux analyses conducted 0, 1, 3, and 5 d after the treatment confirmed this adaptation pattern of NH4+ uptake. Half-lives for NH4+ exchange with subcellular compartments, cytoplasmic NH4+ concentrations, and efflux (as percentage of influx) were unaffected by hypoxia. However, significant differences were observed in the relative amounts of N allocated to NH4+ assimilation and the vacuole versus translocation to the shoot. Kinetic experiments conducted at 100, 50, 35, and 15% O2 saturation showed no significant change in the Km value for NH4+ uptake with varying O2 supply. However, Vmax was 42% higher than controls at 50% O2 saturation, unchanged at 35%, and 10% lower than controls at 15% O2. The significance of these flux adaptations is discussed.  相似文献   

6.
The quadruplex structures of the human telomere sequences AG3(T2AG3)3 I and (T2AG3)4 II were investigated in the presence of Na+ and K+ ions, through the cross-linking of adenines and guanines by the cis- and trans-[Pt(NH3)2(H2O)2](NO3)2 complexes 1 and 2. The bases involved in chelation of the cis- and trans-Pt(NH3)2 moieties were identified by chemical and 3′-exonuclease digestions of the products isolated after denaturing gel electrophoresis. These are the four adenines of each sequence and four out of the 12 guanines. Two largely different structures have been reported for I: A from NMR data in Na+ solution and B from X-ray data of a K+-containing crystal. Structure A alone agrees with our conclusions about the formation of the A1–G10, A13–G22, A1–A13 platinum chelates at the top of the quadruplex and A7–A19, G4–A19 and A7–G20 at the bottom, whether the Na+ or K+ ion is present. At variance with a recent proposal that structures A and B could be the major species in Na+ and K+ solutions, respectively, our results suggest that structure A exists predominantly in the presence of both ions. They also suggest that covalent platinum cross-linking of a human telomere sequence could be used to inhibit telomerase.  相似文献   

7.
The potassium conductance of the basolateral membrane (BLM) of proximal tubule cells is a critical regulator of transport since it is the major determinant of the negative cell membrane potential and is necessary for pump-leak coupling to the Na+,K+-ATPase pump. Despite this pivotal physiological role, the properties of this conductance have been incompletely characterized, in part due to difficulty gaining access to the BLM. We have investigated the properties of this BLM K+ conductance in dissociated, polarized Ambystoma proximal tubule cells. Nearly all seals made on Ambystoma cells contained inward rectifier K+ channels (γslope, in = 24.5 ± 0.6 pS, γchord, out = 3.7 ± 0.4 pS). The rectification is mediated in part by internal Mg2+. The open probability of the channel increases modestly with hyperpolarization. The inward conducting properties are described by a saturating binding–unbinding model. The channel conducts Tl+ and K+, but there is no significant conductance for Na+, Rb+, Cs+, Li+, NH4+, or Cl. The channel is inhibited by barium and the sulfonylurea agent glibenclamide, but not by tetraethylammonium. Channel rundown typically occurs in the absence of ATP, but cytosolic addition of 0.2 mM ATP (or any hydrolyzable nucleoside triphosphate) sustains channel activity indefinitely. Phosphorylation processes alone fail to sustain channel activity. Higher doses of ATP (or other nucleoside triphosphates) reversibly inhibit the channel. The K+ channel opener diazoxide opens the channel in the presence of 0.2 mM ATP, but does not alleviate the inhibition of millimolar doses of ATP. We conclude that this K+ channel is the major ATP-sensitive basolateral K+ conductance in the proximal tubule.  相似文献   

8.
The conserved family of AMT/Rh proteins facilitates ammonium transport across animal, plant, and microbial membranes. A bacterial homologue, AmtB, forms a channel-like structure and appears to function as an NH3 gas channel. To evaluate the function of eukaryotic homologues, the human RhCG glycoprotein and the tomato plant ammonium transporter LeAMT1;2 were expressed and compared in Xenopus oocytes and yeast. RhCG mediated the electroneutral transport of methylammonium (MeA), which saturated with Km = 3.8 mM at pHo 7.5. Uptake was strongly favored by increasing the pHo and was inhibited by ammonium. Ammonium induced rapid cytosolic alkalinization in RhCG-expressing oocytes. Additionally, RhCG expression was associated with an alkali-cation conductance, which was not significantly permeable to NH4+ and was apparently uncoupled from the ammonium transport. In contrast, expression of the homologous LeAMT1;2 induced pHo-independent MeA+ uptake and specific NH4+ and MeA+ currents that were distinct from endogenous currents. The different mechanisms of transport, including the RhCG-associated alkali-cation conductance, were verified by heterologous expression in appropriate yeast strains. Thus, homologous AMT/Rh-type proteins function in a distinct manner; while LeAMT1;2 carries specifically NH4+, or cotransports NH3/H+, RhCG mediates electroneutral NH3 transport.  相似文献   

9.
(6R)-2,2,6-Trimethyl-1,4-cyclohexanedione (levodione) reductase was isolated from a cell extract of the soil isolate Corynebacterium aquaticum M-13. This enzyme catalyzed regio- and stereoselective reduction of levodione to (4R,6R)-4-hydroxy-2,2,6-trimethylcyclohexanone (actinol). The relative molecular mass of the enzyme was estimated to be 142,000 Da by high-performance gel permeation chromatography and 36,000 Da by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The enzyme required NAD+ or NADH as a cofactor, and it catalyzed reversible oxidoreduction between actinol and levodione. The enzyme was highly activated by monovalent cations, such as K+, Na+, and NH4+. The NH2-terminal and partial amino acid sequences of the enzyme showed that it belongs to the short-chain alcohol dehydrogenase/reductase family. This is the first report of levodione reductase.  相似文献   

10.
The bulge region of HIV-1 TAR RNA binds metal ions in solution   总被引:4,自引:1,他引:3       下载免费PDF全文
Binding of Mg2+, Ca2+ and Co(NH3)63+ ions to the HIV-1 TAR RNA in solution was analysed by 19F NMR spectroscopy, metal ion-induced RNA cleavages and Brownian dynamics (BD) simulations. Chemically synthesised 29mer oligoribonucleotides of the TAR sequence labelled with 5-fluorouridine (FU) were used for 19F NMR-monitored metal ion titration. The chemical shift changes of fluorine resonances FU-23, FU-25 and FU-40 upon titration with Mg2+ and Ca2+ ions indicated specific, although weak, binding at the bulge region with the dissociation constants (Kd) of 0.9 ± 0.6 and 2.7 ± 1.7 mM, respectively. Argininamide, inducing largest 19F chemical shifts changes at FU-23, was used as a reference ligand (Kd = 0.3 ± 0.1 mM). In the Pb2+-induced TAR RNA cleavage experiment, strong and selective cleavage of the C24-U25 phosphodiester bond was observed, while Mg2+ and Ca2+ induced cuts at all 3-nt residues of the bulge. The inhibition of Pb2+-specific TAR cleavage by di- and trivalent metal ions revealed a binding specificity [in the order Co(NH3)63+ > Mg2+ > Ca2+] at the bulge site. A BD simulation search of potential magnesium ion sites within the NMR structure of HIV-1 TAR RNA was conducted on a set of 20 conformers (PDB code 1ANR). For most cases, the bulge region was targeted by magnesium cations.  相似文献   

11.
By examining the consequences both of changes of [K+]o and of point mutations in the outer pore mouth, our goal was to determine if the mechanism of the block of Kv1.5 ionic currents by external Ni2+ is similar to that for proton block. Ni2+ block is inhibited by increasing [K+]o, by mutating a histidine residue in the pore turret (H463Q) or by mutating a residue near the pore mouth (R487V) that is the homolog of Shaker T449. Aside from a slight rightward shift of the Q-V curve, Ni2+ had no effect on gating currents. We propose that, as with Ho+, Ni2+ binding to H463 facilitates an outer pore inactivation process that is antagonized by Ko+ and that requires R487. However, whereas Ho+ substantially accelerates inactivation of residual currents, Ni2+ is much less potent, indicating incomplete overlap of the profiles of these two metal ions. Analyses with Co2+ and Mn2+, together with previous results, indicate that for the first-row transition metals the rank order for the inhibition of Kv1.5 in 0 mM Ko+ is Zn2+ (KD ~ 0.07 mM) ≥ Ni2+ (KD ~ 0.15 mM) > Co2+ (KD ~ 1.4 mM) > Mn2+ (KD > 10 mM).  相似文献   

12.
Source and Magnitude of Ammonium Generation in Maize Roots   总被引:1,自引:0,他引:1       下载免费PDF全文
Studies with 15N indicate that appreciable generation of NH4+ from endogenous sources accompanies the uptake and assimilation of exogenous NH4+ by roots. To identify the source of NH4+ generation, maize (Zea mays L.) seedlings were grown on 14NH4+ and then exposed for 3 d to highly labeled 15NH4+. More of the entering 15NH4+ was incorporated into the protein-N fraction of roots in darkness (approximately 25%) than in the light (approximately 14%). Although the 14NH4+ content of roots declined rapidly to less than 1 μmol per plant, efflux of 14NH4+ continued throughout the 3-d period at an average daily rate of 14 μmol per plant. As a consequence, cumulative 14NH4+ efflux during the 3-d period accounted for 25% of the total 14N initially present in the root. Although soluble organic 14N in roots declined during the 3-d period, insoluble 14N remained relatively constant. In shoots both soluble organic 14N and 14NH4+ declined, but a comparable increase in insoluble 14N was noted. Thus, total 14N in shoots remained constant, reflecting little or no net redistribution of 14N between shoots and roots. Collectively, these observations reveal that catabolism of soluble organic N, not protein N, is the primary source of endogenous NH4+ generation in maize roots.  相似文献   

13.
HCN channels are thought to be structurally similar to Kv channels, but show much lower selectivity for K+. The ∼3.3 Å selectivity filter of K+ channels is formed by the pore-lining sequence XT(V/I)GYG, with X usually T, and is held stable by key residues in the P-loop. Differences in the P-loop sequence of HCN channels (eg. the pore-lining sequence L478C479IGYG) suggest these residues could account for differences in selectivity between these channel families. Despite being expressed, L478T/C479T HCN4 channels did not produce current. Since threonine in the second position is highly conserved in K+ channels, we also studied C479T channels. Based on permeability ratios (PX/PK), C479T HCN4 channels (K+(1)>Rb+(0.85)>Cs+(0.59)>Li+(0.50)≥Na+(0.49)) were less selective than WT rabbit HCN4 (K+(1)>Rb+(0.48)>Cs+(0.31)≥Na+(0.29)>Li+(0.03)), indicating that the TIGYG sequence is insufficient to confer K+ selectivity to HCN channels. C479T HCN4 channels had an increased permeability to large organic cations than WT HCN4 channels, as well as increased unitary K+ conductance, and altered channel gating. Collectively, these results suggest that HCN4 channels have larger pores than K+ channels and replacement of the cysteine at position 479 with threonine further increases pore size. Furthermore, selected mutations in other regions linked previously to pore stability in K+ channels (ie. S475D, S475E and F471W/K472W) were also unable to confer K+ selectivity to C479T HCN4 channels. Our findings establish the presence of the TIGYG pore-lining sequence does not confer K+ selectivity to rabbit HCN4 channels, and suggests that differences in selectivity of HCN4 versus K+ channels originate from differences outside the P-loop region.  相似文献   

14.
The potentially toxic freshwater cyanobacterium Cylindrospermopsis raciborskii has become increasingly prevalent in tropical and temperate water bodies worldwide. This paper investigates the effects of different nitrogen sources (NO3, NH4+, and omission of a fixed form of nitrogen) on the growth rates, morphologies, and cylindrospermopsin (CYL) concentrations (expressed as a percentage of the freeze-dried weight) of seven C. raciborskii isolates obtained from a range of water bodies in northern Australia and grown in batch culture. In general, growth rates were lowest in the absence of a fixed-nitrogen source and highest with NH4+ as the nitrogen source. Conversely, the highest concentrations of CYL were recorded in cultures grown in the absence of a fixed-nitrogen source and the lowest were found in cultures supplied with NH4+. Cultures supplied with NO3 were intermediate with respect to both CYL concentration and growth rate. Different nitrogen sources resulted in significant differences in the morphology of C. raciborskii trichomes. Most notable were the loss of heterocysts and the tapering of end cells in cultures supplied with NH4+ and the statistically significant increase in vegetative cell length (nitrogen depleted < NO3 < NH4+). The morphological changes induced by different nitrogen sources were consistent for all isolates, despite measurable differences in vegetative-cell and heterocyst dimensions among isolates. Such induced morphological variation has implications for Cylindrospermopsis taxonomy, given that distinctions between species are based on minor and overlapping differences in cell lengths and widths. The close phylogenetic association among all seven isolates was confirmed by the high level (>99.8%) of similarity of their 16S rRNA gene sequences. Another genetic technique, analysis of the HIP1 octameric-palindrome repeated sequence, showed greater heterogeneity among the isolates and appears to be a useful method for distinguishing among isolates of C. raciborskii.  相似文献   

15.
Four genomic DNAs of differing GC content (Micrococcus luteus, 72% GC; Escherichia coli, 50% GC; calf thymus, 42% GC; Clostridium perfringens, 27% GC) have been employed as targets of interaction by the cationic polyamines spermidine {[H3N(CH2)3NH2(CH2)4NH3]3+} and spermine {[(CH2)4(NH2(CH2)3NH3)2]4+}. In solutions containing 60 mM DNA phosphate (~20 mg DNA/ml) and either 1, 5 or 60 mM polyamine, only Raman bands associated with the phosphates exhibit large spectral changes, demonstrating that B-DNA phosphates are the primary targets of interaction. Phosphate perturbations, which are independent of base composition, are consistent with a model of non-specific cation binding in which delocalized polyamines diffuse along DNA while confined by the strong electrostatic potential gradient perpendicular to the helix axis. This finding provides experimental support for models in which polyamine-induced DNA condensation is driven by non-specific electrostatic binding. The Raman spectra also demonstrate that major groove sites (guanine N7 and thymine C5H3) are less affected than phosphates by polyamine–DNA interactions. Modest dependence of polyamine binding on genome base composition suggests that sequence context plays only a secondary role in recognition. Importantly, the results demonstrate that polyamine binding has a negligible effect on the native B-form secondary structure. The capability of spermidine or spermine to bind and condense genomic B-DNA without disrupting the native structure must be taken into account when considering DNA organization within bacterial nucleoids or cell nuclei.  相似文献   

16.
Nitrate-ammonium synergism in rice. A subcellular flux analysis   总被引:12,自引:0,他引:12       下载免费PDF全文
Many reports have shown that plant growth and yield is superior on mixtures of NO3 and NH4+ compared with provision of either N source alone. Despite its clear practical importance, the nature of this N-source synergism at the cellular level is poorly understood. In the present study we have used the technique of compartmental analysis by efflux and the radiotracer 13N to measure cellular turnover kinetics, patterns of flux partitioning, and cytosolic pool sizes of both NO3 and NH4+ in seedling roots of rice (Oryza sativa L. cv IR72), supplied simultaneously with the two N sources. We show that plasma membrane fluxes for NH4+, cytosolic NH4+ accumulation, and NH4+ metabolism are enhanced by the presence of NO3, whereas NO3 fluxes, accumulation, and metabolism are strongly repressed by NH4+. However, net N acquisition and N translocation to the shoot with dual N-source provision are substantially larger than when NO3 or NH4+ is provided alone at identical N concentrations.  相似文献   

17.
Enterococcus hirae ATCC 9790 grew well in Na+-deficient, low-K+ medium, but growth was inhibited by carbonylcyanide m-chlorophenylhydrazone (CCCP). Growth inhibition and decrease of cellular K+ levels in the presence of CCCP were relieved by the addition of Na+ and a high concentration of K+. In contrast, in the mutant defective in Na+-ATPase or the NtpJ component of the KtrII K+ uptake system, CCCP-induced growth inhibition was rescued by a high concentration of K+ but not of Na+. These transporters are thus indispensable for homeostatis of K+ and Na+ at low proton potential.  相似文献   

18.
A bacterium, Ochrobactrum anthropi, produced a large amount of a nucleosidase when cultivated with purine nucleosides. The nucleosidase was purified to homogeneity. The enzyme has a molecular weight of about 170,000 and consists of four identical subunits. It specifically catalyzes the irreversible N-riboside hydrolysis of purine nucleosides, the Km values being 11.8 to 56.3 μM. The optimal activity temperature and pH were 50°C and pH 4.5 to 6.5, respectively. Pyrimidine nucleosides, purine and pyrimidine nucleotides, NAD, NADP, and nicotinamide mononucleotide are not hydrolyzed by the enzyme. The purine nucleoside hydrolyzing activity of the enzyme was inhibited (mixed inhibition) by pyrimidine nucleosides, with Ki and Ki′ values of 0.455 to 11.2 μM. Metal ion chelators inhibited activity, and the addition of Zn2+ or Co2+ restored activity. A 1.5-kb DNA fragment, which contains the open reading frame encoding the nucleosidase, was cloned, sequenced, and expressed in Escherichia coli. The deduced 363-amino-acid sequence including a 22-residue leader peptide is in agreement with the enzyme molecular mass and the amino acid sequences of NH2-terminal and internal peptides, and the enzyme is homologous to known nucleosidases from protozoan parasites. The amino acid residues forming the catalytic site and involved in binding with metal ions are well conserved in these nucleosidases.  相似文献   

19.
Additions of ammonium and non-ammonium salts inhibit atmospheric methane consumption by soil at salt concentrations that do not significantly affect the soil water potential. The response of soils to non-ammonium salts has previously raised questions about the mechanism of ammonium inhibition. Results presented here show that inhibition of methane consumption by non-ammonium salts can be explained in part by ion-exchange reactions: cations desorb ammonium, with the level of desorption varying as a function of both the cation and anion added; differential desorption results in differential inhibition levels. Differences in the extent of inhibition among ammonium salts can also be explained in part by the effects of anions on ammonium exchange. In contrast, only minimal effects of cations and anions are observed in liquid cultures of Methylosinus trichosporium OB3b. The comparable level of inhibition by equinormal concentrations of NH4Cl and (NH4)2SO4 and the insensitivity of salt inhibition to increasing methane concentrations (from 10 to 100 ppm) are of particular interest, since both of these patterns are in contrast to results for soils. The greater inhibition of methane consumption for NH4Cl than (NH4)2SO4 in soils can be attributed to increased ammonium adsorption by sulfate; increasing inhibition by non-ammonium salts with increasing methane concentrations can be attributed to desorbed ammonium and a physiological mechanism proposed previously for pure cultures.  相似文献   

20.
The solution structures of two 27 nt RNA hairpins and their complexes with cobalt(III)-hexammine [Co(NH3)63+] were determined by NMR spectroscopy. The RNA hairpins are variants of the P4 region from Escherichia coli RNase P RNA: a U-to-A mutant changing the identity of the bulged nucleotide, and a U-to-C, C-to-U double mutant changing only the bulge position. Structures calculated from NMR constraints show that the RNA hairpins adopt different conformations. In the U-to-C, C-to-U double mutant, the conserved bulged uridine in the P4 wild-type stem is found to be shifted in the 3′-direction by one nucleotide when compared with the wild-type structure. Co(NH3)63+ is used as a spectroscopic probe for Mg(H2O)62+ binding sites because both complexes have octahedral symmetry and have similar radii. Intermolecular NOE crosspeaks between Co(NH3)63+ and RNA protons were used to locate the site of Co(NH3)63+ binding to both RNA hairpins. The metal ion binds in the major groove near a bulge loop in both mutants, but is shifted 3′ by about one base pair in the double mutant. The change of the metal ion binding site is compared with results obtained on corresponding mutant RNase P RNA molecules as reported by Harris and co-workers (RNA, 1, 210–218).  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