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 15 毫秒
1.
Electron paramagnetic resonance (epr) studies demonstrate that at low levels of conalbumin (CA) saturation with Fe3+ or VO2+, a ph-dependent preference of the metal exists for different protein binding-site configurations,A, B, and C. The vanadyl ion epr spectra of mixed VO2+, Fe3+-conalbumin in which Fe3+ is preferentially bound to the N- or C-terminal binding site are consistent with all three configurations being formed at both metal sites. At high pH the spectra suggest interaction between binding sites. In the absence of HCO3?, VO2+ is bound almost exclusively in B configuration; a full binding capacity of 2 VO2+ per CA is retained. Stoichiometric amounts of HCO3? convert the epr spectrum from B to an A, B, C type. Addition of oxalate to bicarbonate-free preparations converts the B spectrum to an A′, B, C′ type where the B resonances have lost intensity to the A′ and C′ resonances but have not changed position. The data suggest that configuration B is anion independent and that only one equivalent of binding sites at pH 9 responds to the presence of HCO31? or oxalate by changing configuration but not metal binding capability. The form of the bound anion may be HCO3? rather than CO32?. The formation rate of the colored ferric conalbumin complex by oxidizing Fe2+ to Fe3+ in limited HCO3? at pH 9 is also consistent with one equivalent of sites having different anion requirements than the remaining sites. Increased NaCl or NaClO4 concentration or substitution of D2O for water as solvent affect the environment of bound VO2+, but the mechanisms of action are unknown.  相似文献   

2.
A laccase from the culture filtrate of Phellinus linteus MTCC-1175 has been purified to homogeneity. The method involved concentration of the culture filtrate by ammonium sulphate precipitation and an anion exchange chromatography on DEAE-cellulose. The SDS-PAGE and native-PAGE gave single protein band indicating that the enzyme preparation was pure. The molecular mass of the enzyme determined from SDS-PAGE analysis was 70 kDa. Using 2.6-dimethoxyphenol, 2.2′[azino-bis-(3-ethylbonzthiazoline-6-sulphonic acid) diammonium salt] (ABTS) and 4-hydroxy-3,5-dimethoxybenzaldehyde azine as the substrates, the K m, k cat and k cat/K m values of the laccase were found to be 160 μM, 6.85 s?1, 4.28 × 104 M?1 s?1, 42 μM, 6.85 s?1, 16.3 × 104 M?1 s?1 and 92 μM, 6.85 s?1, 7.44 × 104 M?1 s?1, respectively. The pH and the temperature optima of the P. linteus MTCC-1175 laccase were 5.0 and 45°C, respectively. The activation energy for thermal denaturation of the enzyme was 38.20 kJ/mole/K. The enzyme was the most stable at pH 5.0 after 1 h reaction. In the presence of ABTS as the mediator, the enzyme transformed toluene, 3-nitrotoluene and 4-chlorotoluene to benzaldehyde, 3-nitrobenzaldehyde and 4-chlorobenzaldehyde, respectively.  相似文献   

3.
We studied anionic inhibition of the reaction CO2 + OH?? HCO3? catalyzed by human red cell carbonic anhydrase B (I) and C (II), using iodide and cyanate. In the forward reaction with respect to CO2 as the substrate, inhibition was mixed but favoring noncompetitive; the back reaction, with HCO3? as the substrate, yielded strict competitive kinetics. Mean inhibition constants, KI, in the pH range 7.2–7.5 are: iodide, 0.5 mm for enzyme B and 16 mm for C; cyanate, 0.8 μm for B and 20 μm for C. When OH? was considered as the substrate for the forward reaction, cyanate and chloride behaved as competitive inhibitors. The true inhibition constant (KI0) for cyanate (calculated for infinitely low OH?) is 0.4 μm for enzyme B and 4 μm for C. Apart from the difference in anion affinity and some 10-fold higher activity of C > B, the isozymes showed similar patterns of inhibition. Data agree with generally proposed mechanisms describing the active site as ZnH2O with pKa of about 7.  相似文献   

4.
An acidic polygalacturonase (PG) secreted by Rhizopus oryzae MTCC-1987 in submerged fermentation condition has been purified to electrophoretic homogeneity using ammonium sulphate fractionation and anion exchange chromatography on diethylaminoethyl cellulose. The purified enzyme gave a single protein band in sodium dodecyl sulphatepolyacrylamide gel electrophoresis analysis with a molecular mass corresponding to 75.5 kDa. The K m and k cat values of the PG were 2.7 mg/mL and 2.23 × 103 s?1, respectively, using citrus polygalacturonic acid as the substrate. The optimum pH of the purified PG was 5.0 and it does not loose activity appreciably if left for 24 hours in the pH range from 5.0 to 12.0. The optimum temperature of purified enzyme was 50°C and the enzyme does not loose activity below 30°C if exposed for two hours. The purified enzyme showed complete inhibition with 1 mM Ag+, Hg2+ and KMnO4, while it was stimulated to some extent by Co2+. The purified PG exhibited retting of Crotalaria juncea fibre in absence of ethylenediaminetetraacetic acid.  相似文献   

5.
Two l-lactate dehydrogenase isoenzymes and one dl-lactate dehydrogenase could be separated from potato tubers by polyacrylamide-gel electrophoresis. The enzymes are specific for lactate, while β-hydroxybutyric acid, glycolic acid, and glyoxylic acid are not oxidized. Their pH optima are pH 6.9 for the oxidation and 8.0 for the reduction reaction.The Km values for l-lactate for the two isoenzymes are 2.00 × 10?2 and 1.82 × 10?2, m. In the reverse reaction the affinities for pyruvate are 3.24 × 10?4 and 3.34 × 10?4, m. Both enzymes have similar affinities for NAD and NADH (3.00 × 10?4; 4.00 × 10?4, and 8.35 × 10?4; 5.25 × 10?4, m).The dl-lactate oxidoreductase may transfer electrons either to NAD or N-methyl-phenazinemethosulfate. The Km values of this enzyme for l-lactate are 4.5 × 10?2, m and for d-lactate 3.34 × 10?2, m. Its affinity for pyruvate is 4.75 × 10?4, m. The enzyme is inhibited by excess NAD (Km = 1.54 × 10?4, M) and has an affinity toward NADH (Km = 5.00 × 10?3, M) which is about one tenth of that of the two isoenzymes of l-lactate dehydrogenase.  相似文献   

6.
Adaptive capacity of bacteria and archaea from salt lakes of the Altai Region are discussed. It is established that halophilic archaea (genus Halorubrum) and halotolerant bacteria (genus Halomonas) grow in a wide range of pH and mineralization (in the presence of Cl?, SO 4 2? , ClO 4 ? , Mg2+) and survive at low temperatures with a minor decrease in viability.  相似文献   

7.
Activation by different anions of γ-glutamyltransferase obtained in a. particulate form from fruiting bodies of Lentinus edodes has been studied using either L-γ-glutamyl-p-nitroanlide or lentinic acid as substrate. The mushroom transferase was activated by SCN?, NO3?, Cl?, Br?, ClO3?, Bro3?, N3?, I? and F?, but not those alkali and earth cations previously believed to activate the animal transferase, nor by citrate, claimed to be effective for the kidney bean transferase. Among anions proved hardly to activate the transferase were ClO4?, NO2?, HCO3?, H2PO4?, SO32? and SO42?. A high concentration of these anions more or less impeded the halide activation. Kinetic studies revealed that halides function as activators of increasing Vmax while keeping Km constant. These observations appeared least compatible with the possibility that the anion activation might involve a non-specific effect of high solute concentration, viz. dissociation of the enzyme from the supporting structure in the particulates. The activating effect of halides described here probably extends also to the animal enzymes.  相似文献   

8.
C.A. Wraight 《BBA》1979,548(2):309-327
The photoreduction of ubiquinone in the electron acceptor complex (Q1Q11) of photosynthetic reaction centers from Rhodopseudomonas sphaeroides, R26, was studied in a series of short, saturating flashes. The specific involvement of H+ in the reduction was revealed by the pH dependence of the electron transfer events and by net H+ binding during the formation of ubiquinol, which requires two turnovers of the photochemical act. On the first flash Q11 receives an electron via Q1 to form a stable ubisemiquinone anion (Q??11); the second flash generates Q??1. At low pH the two semiquinones rapidly disproportionate with the uptake of 2 H+, to produce Q11H2. This yields out-of-phase binary oscillations for the formation of anionic semiquinone and for H+ uptake. Above pH 6 there is a progressive increase in H+ binding on the first flash and an equivalent decrease in binding on the second flash until, at about pH 9.5, the extent of H+ binding is the same on all flashes. The semiquinone oscillations, however, are undiminished up to pH 9. It is suggested that a non-chromophoric, acid-base group undergoes a pK shift in response to the appearance of the anionic semiquinone and that this group is the site of protonation on the first flash. The acid-base group, which may be in the reaction center protein, appears to be subsequently involved in the protonation events leading to fully reduced ubiquinol. The other proton in the two electron reduction of ubiquinone is always taken up on the second flash and is bound directly to Q??11. At pH values above 8.0, it is rate limiting for the disproportionation and the kinetics, which are diffusion controlled, are properly responsive to the prevailing pH. Below pH 8, however, a further step in the reaction mechanism was shown to be rate limiting for both H+ binding electron transfer following the second flash.  相似文献   

9.
E.J. Land  A.J. Swallow 《BBA》1974,368(1):86-96
When ferricytochrome c at pH about 9 is reduced by hydrated electrons and/or CO2?, it gives rise to an unstable form of ferrocytochrome c whose absorption spectrum, particularly in the Soret region, differs from that of normal ferrocytochrome c. This form changes intramolecularly (life-time about 0.1 s at ambient temperature) to yield normal ferrocytochrome c, and by 0.5 s the change in absorption spectrum in the range 225–600 nm produced by e?aq and/or CO2? is identical to the final change produced by reduction with an equivalent amount of sodium dithionite. This shows that both e?aq and CO?2 reduce cytochrome c with practically 100% efficiency. In the range 600–800 nm the spectrum of the unstable form is the same as that of normal ferrocytochrome c, both having small absorptions at 695 nm as compared with ferricytochrome c. As the unstable form disappears however a further loss of absorption at 695 nm occurs. This is taken to imply that the unstable form decays to a second unstable form which then rapidly donates an electron to the unchanged neutral form of ferricytochrome c, so reducing absorption in the 695 nm band. Subsequent to this process the absorption in the 695 nm band increases over a period of minutes owing to re-equilibration between the neutral and alkaline formes of ferricytochrome c. Between pH 7 and 10 the effect of pH on the absorption changes is consistent with the hypothesis of a second unstable form of ferrocytochrome c. Additional phenomena arise in more alkaline solutions. The rates of the various unimolecular processes are thought to be determined by the rates of change of conformation of the protein parts of the molecule following the change in oxidation state.  相似文献   

10.
A putative endo-1,4-β-d-xylanohydrolase gene xyl11 from Aspergillus niger, encoding a 188-residue xylanase of glycosyl hydrolase family 11, was constitutively expressed in Pichia pastoris. The recombinant Xyl11 exhibited optimal activity at pH 5.0 and 50 °C, and displayed more than 68 % of the maximum activity over the temperature range 35–65 °C and 33 % over the pH range 2.2–7.0. It maintained more than 40 % of the original activity after incubation at 90 °C (pH 5.0) for 10 min and more than 75 % of the original activity after incubation at pH 2.2–11.0 (room temperature) for 2 h. The specific activity, K m and V max of purified Xyl11 were 22,253 U mg?1, 6.57 mg ml?1 and 51,546.4 μmol min?1 mg?1. It could degrade xylan to a series of xylooligosaccharides and no xylose was detected. The recombinant enzyme with high stability and catalytic efficiency could work over wide ranges of pH and temperature and thus has the potential for various industrial applications.  相似文献   

11.
The kinetics of the binding of cyanide to ferric chloroperoxidase have been studied at 25°C and ionic strength 0.11 M using a stopped-flow apparatus. The dissociation constant (KCN) of the peroxidase-cyanide complex and both forward (k+) and reverse (k?) rate constants are independent of the H+ concentration over the pH range 2.7 to 7.1. The values obtained are kcn = (9.5 ± 1.0) × 10-5 M, k+. = (5.2 ± 0.5) × 104 M?1 sec?1 and k- = (5.0± 1.4) sec-1. In the presence of 0 06 M potassium nitrate the affinity of cyanide for chloroperoxidase decreases due to the inhibition of the forward reaction. The dissociation rate is not affected. The nitrate anion exerts its influence by binding to a protonated form of the enzyme, whereas the cyanide binds to the unprotonated form. Binding of nitrate results in an apparent shift towards higher pKa values of the ionization of a crucial heme-linked acid group. Hence the influence of this group can be detected in the accessible pH range. Extrapolation to zero nitrate concentration yields a value of 3.1±0.3 for the pKa of the heme-linked acid group.  相似文献   

12.
Heavy metal ions (Pb2+, Cd2+, Mn2+, Cu2+, and Cr2O7 2?) were biosorbed by brown seaweeds (Hizikia fusiformis, Laminaria japonica, and Undaria pinnatifida) collected from the southern coast of South Korea. The biosorption of heavy metal ions was pH-dependent showing a minimum absorption at pH 2 and a maximum biosorption at pH 4 (Pb2+, Cd2+, Mn2+, and Cr2O7 2?) or pH 6 (Cu2+). Biosorption increased most noticeably for pH changes from 2 to 3. In the latter pH range, biosorption increased, because a higher pH decreased the electrostatic repulsion between metal ions and functional groups on the seaweed. In the pH range of 2 ~ 4, biosorption of negatively-charged chromium species (Cr2O7 ?2) followed the pattern of positively-charged metal ions (Pb2+, Cd2+, Mn2+, and Cu2+). This suggests that the most prevalent chromium species were positively-charged Cr3+, reduced from Cr6+ in Cr2O7 ?2. Whereas positively-charged heavy metal ions (Pb2+, Cd2+, Mn2+, and Cu2+) reached a plateau after the maximum level, biosorption of chromium ions decreased noticeably between pH 5 and 8. Kinetic data showed that biosorption by brown seaweed occurred rapidly during the first 10 min, and most of the heavy metals were bound to the seaweed within 30 min. Equilibrium adsorption data for a lead ion could fit well in the Langmuir and Freundlich isotherm models with regression coefficients (R 2) between 0.93 and 0.98.  相似文献   

13.
The permeability of phospholipid membranes to the superoxide anion (O2?) was determined using soybean phospholipid vesicles containing FMN in the internal space. The efflux of O2? generated by the illumination of FMN was so slow that more than 90% of the radicals were spontaneously disproportionated within the vesicles before they could react with cytochrome c at the membrane exterior. The amount of diffused O2? was proportional to the intravesicular concentration of O2? over a range from 1 to 10 μm which was deduced from its disproportionation rate. The permeability coefficient of the phospholipid bilayer for O2? was estimated to be 2.1 × 10?6 cm s?1 at pH 7.3 and 25 ° C. Superoxide dismutase trapped inside vesicles was not reactive with extravesicular O2? unless Triton X-100 was added. O2? generated outside spinach chloroplast thylakoids did not interact with superoxide dismutase or cytochrome c which had been enclosed in the thylakoids. Thus, chloroplast thylakoids also showed little permeability to O2?.  相似文献   

14.
There has been considerable interest in cultivation of green microalgae (Chlorophyta) as a source of lipid that can alternatively be converted to biodiesel. However, almost all mass cultures of algae are carbon-limited. Therefore, to reach a high biomass and oil productivities, the ideal selected microalgae will most likely need a source of inorganic carbon. Here, growth and lipid productivities of Tetraselmis suecica CS-187 and Chlorella sp were tested under various ranges of pH and different sources of inorganic carbon (untreated flue gas from coal-fired power plant, pure industrial CO2, pH-adjusted using HCl and sodium bicarbonate). Biomass and lipid productivities were highest at pH 7.5 (320?±?29.9 mg biomass L?1 day?1and 92?±?13.1 mg lipid L?1 day?1) and pH 7 (407?±?5.5 mg biomass L?1 day?1 and 99?±?17.2 mg lipid L?1 day?1) for T. suecica CS-187 and Chlorella sp, respectively. In general, biomass and lipid productivities were pH 7.5?>?pH 7?>?pH 8?>?pH 6.5 and pH 7?>?pH 7.5?=?pH 8?>?pH 6.5?>?pH 6?>?pH 5.5 for T. suecica CS-187 and Chlorella sp, respectively. The effect of various inorganic carbon on growth and productivities of T. suecica (regulated at pH?=?7.5) and Chlorella sp (regulated at pH?=?7) grown in bag photobioreactors was also examined outdoor at the International Power Hazelwood, Gippsland, Victoria, Australia. The highest biomass and lipid productivities of T. suecica (51.45?±?2.67 mg biomass L?1 day?1 and 14.8?±?2.46 mg lipid L?1 day?1) and Chlorella sp (60.00?±?2.4 mg biomass L?1 day?1 and 13.70?±?1.35 mg lipid L?1 day?1) were achieved when grown using CO2 as inorganic carbon source. No significant differences were found between CO2 and flue gas biomass and lipid productivities. While grown using CO2 and flue gas, biomass productivities were 10, 13 and 18 %, and 7, 14 and 19 % higher than NaHCO3, HCl and unregulated pH for T. suecica and Chlorella sp, respectively. Addition of inorganic carbon increased specific growth rate and lipid content but reduced biomass yield and cell weight of T. suecica. Addition of inorganic carbon increased yield but did not change specific growth rate, cell weight or content of the cell weight of Chlorella sp. Both strains showed significantly higher maximum quantum yield (Fv/Fm) when grown under optimum pH.  相似文献   

15.
The effect of different co-anions on the formation and aggregation of the ordered structure of the anionic polysaccharide kappa-carrageenan has been investigated by optical rotation, differential scanning calorimetry, and halide-n.m.r. spectroscopy. The mid-point temperature (Tm) of the disorder—order transition increases systematically with the Hofmeister number for the anion through the lyotropic series SO42? < F? < Cl? < Br? < NO3? < I? < SCN? when salt concentration and cation (Me4N+ or K+) are held constant. A corresponding increase is observed in transition enthalpy (ΔHcal) and entropy (ΔScal). Helix—helix aggregation (as indicated by turbidity, gel formation, and hysteresis between heating and cooling scans) also shows a systematic dependence on the Hofmeister number for the anion, but in the opposite sense. Thus, with tetramethylammonium as the sole counterion present, clear solutions with no thermal hysteresis in the order—disorder transition are observed at all temperatures with I?, Br?, NO3?, or Cl? as co-anion, whereas weak, turbid gels with significant thermal hysteresis between melting and setting are formed in the presence of SO42?, and to a lesser degree F?. With K+ as counterion, a similar regular progression is observed through the anion lyotropic series from rapid formation of very turbid gels in the presence of F?, to very slow development of clear gels with I? or SCN?. In agreement with previous studies, an increase in 127I-n.m.r. linewidth was observed on conformational ordering of kappa-carrageenan (Me4N+ salt form) in the presence of Me4NI. However, closely similar behaviour was observed for 35Cl and 81Br, indicating a simple charge-cloud interaction rather than the specific site-binding of I? which has previously been suggested.  相似文献   

16.
The purified extracellular xylanase of polyextremophilic Bacillus halodurans TSEV1 has been visualized as a single band on SDS-PAGE and eluted as single peak by gel filtration, with a molecular mass of 40 kDa. The peptide finger print and cloned xylanase gene sequence analyses indicate that this enzyme belongs to GH family 10. The active site carboxyl residues are mainly involved in catalysis, while tryptophan residues are involved in substrate binding. The enzyme is optimally active at 80 °C and pH 9.0, and stable in the pH range of 7.0–12.0 with T 1/2 of 35 min at 80 °C (pH 9.0). Activation energy for birch wood xylan hydrolysis is 30.51 kJ mol?1. The K m, V max and k cat (birchwood xylan) are 2.05 mg ml?1, 333.33 μmol mg?1 min?1 and 3.33 × 104 min?1, respectively. The pKa1 and pKa2 of ionizable groups of the active site that influence V max are 8.51 and 11.0. The analysis of thermodynamic parameters for xylan hydrolysis suggests this as a spontaneous process. The enzyme is resistant to chemical denaturants like urea and guanidinium-HCl. The site-directed mutagenesis of catalytic glutamic acid residues (E196 and E301) resulted in a complete loss of activity. The birch wood xylan hydrolyzate contained xylobiose and xylotriose as the main products without any trace of xylose, and the enzyme hydrolyzes xylotetraose and xylopentaose rapidly to xylobiose. Thermo-alkali-stability, resistance to various chemical denaturants and mode of action make it a useful biocatalyst for generating xylo-oligosaccharides from agro-residues and bleaching of pulp in paper industries.  相似文献   

17.
The reductant of ferricytochrome c2 in Rhodopseudomonas sphaeroides is a component, Z, which has an equilibrium oxidation-reduction reaction involving two electrons and two protons with a midpoint potential of 155 mV at pH 7. Under energy coupled conditions, the reduction of ferricytochrome c2 by ZH2 is obligatorily coupled to an apparently electrogenic reaction which is monitored by a red shift of the endogeneous carotenoids. Both ferricytochrome c2 reduction and the associated carotenoid bandshift are similarly affected by the concentrations of ZH2 and ferricytochrome c2, pH, temperature the inhibitors diphenylamine and antimycin, and the presence of ubiquinone. The second-order rate constant for ferricytochrome c2 reduction at pH 7.0 and at 24°C was 2 · 109 M?1 · s?1, but this varied with pH, being 5.1 · 108 M?1 · s?1 at pH 5.2 and 4.3 · 109 M?1 · s?1 at pH 9.3. At pH 7 the reaction had an activation energy of 10.3 kcal/mol.  相似文献   

18.
The electronic spectra of NCS? and I? adducts of cobalt(II) human carbonic anhydrase I are pH dependent at pH values below 7. The pKa of such equilibrium is dependent on the anion concentration and varies between 4.6 and 6.6. The 1H NMR spectra show that the three histidine residues are bound to the metal ion over the entire pH range investigated. It is supposed that a Glu residue triggers the change in stereochemistry around the metal ion. It is possible that such a Glu residue is Glu 106 present in the active cavity.  相似文献   

19.
Abstract

Addition of Na2CO3 to almost salt-free DNA solution (5·10?5M EDTA, pH=5.7, Tm=26.5 °C) elevates both pH and the DNA melting temperature (Tm) if Na2CO3 concentration is less than 0.004M. For 0.004M Na2CO3, Tm=58 °C is maximal and pH=10.56. Further increase in concentration gives rise to a monotonous decrease in Tm to 37 °C for 1M N2CO3 (pH=10.57). Increase in pH is also not monotonous. The highest pH=10.87 is reached at 0.04M Na2CO3 (Tm=48.3 °C). To reveal the cause of this DNA destabilization, which happens in a narrow pH interval (10.56÷10.87) and a wide Na2CO3 concentration interval (0.004÷1M), a procedure has been developed for determining the separate influences on Tm of Na+, pH, and anions formed by Na2CO3 (HCO3 ? and CO3 2-). Comparison of influence of anions formed by Na2CO3 on DNA stability with Cl? (anion inert to DNA stability), ClO4 ? (strong DNA destabilizing “chaotropic” anion) and OH? has been carried out. It has been shown that only Na+ and pH influence Tm in Na2CO3 solution at concentrations lower than 0.001M. However, the Tm decrease with concentration for [Na2CO3]≥0.004M is only partly caused by high pH≈10.7. Na2CO3 anions also exert a strong destabilizing influence at these concentrations. For 0.1M Na2CO3 (pH=10.84, [Na+]=0.2M, Tm=42.7 °C), the anion destabilizing effect is higher 20 °C. For NaClO4 (ClO4 ? is a strong “chaotropic” anion), an equal anion effect occurs at much higher concentrations ~3M. This means that Na2CO3 gives rise to a much stronger anion effect than other salts. The effect is pH dependent. It decreases fivefold at neutral pH after addition of HCl to 0.1M Na2CO3 as well as after addition of NaOH for pH>11.2.  相似文献   

20.
Z. Wang  J. Shen  F. Zhang 《Plant and Soil》2006,287(1-2):247-256
The study examined the interactive effect of pH and P supply on cluster-root formation, carboxylate exudation and proton release by an alkaline-tolerant lupin species (Lupinus pilosus Murr.) in nutrient solution. The plants were exposed to 1 (P1, deficient) and 50 μM P (P50, adequate) for 34 days in nutrient solution at either pH 5.6 or 7.8. Plant biomass was not influenced by pH at P1, but at P50 shoot and root dry weights were 23 and 18% higher, respectively, at pH 7.8 than at pH 5.6. There was no significant difference in plant biomass between two P treatments regardless of medium pH. Phosphorus deficiency increased significantly the number of the second-order lateral roots compared with the P50 treatment. Both total root length and specific root length of plants grown at pH 5.6 were higher than those at pH 7.8 regardless of P supply. Cluster roots were formed at P1, but cluster-root number was 2-fold higher at pH 7.8 than pH 5.6. Roots released 16 and 31% more protons at pH 5.6 and 7.8, respectively, in P1 than in P50 treatments, and the rate of proton release followed the similar pattern. At pH 5.6, citrate exudation rate was 0.39 μmol g−1 root DW h−1 at P1, but was under the detection limit at P50; at pH 7.8, it was 2.4-fold higher in P1 than in P50 plants. High pH significantly increased citrate exudation rate in comparison to pH 5.6. The uptake of anions P and S was inhibited at P1 and high pH increased cations Na, Mg and Ca uptake. The results suggested that enhanced cluster-root formation, proton release and citrate exudation may account for the mechanism of efficient P acquisition by alkaline-tolerant L. pilosus well adapted to calcareous soils. Cluster-root formation and citrate exudation in L. pilosus can be altered by medium pH and P deficiency. Phosphorus deficiency-induced proton release may be associated with the reduced anion uptake, but high pH-induced proton release may be partly attributed to increased cation uptake.  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