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 23 毫秒
1.
One mannanase and one of the three xylanases produced by Ceriporiopsis subvermispora grown on Pinus taeda wood chips were characterized. A combination of ion exchange chromatography and SDS-PAGE data revealed the existence of a high-molecular-weight mannanase of 150 kDa that was active against galactoglucomannan and xylan. Its activity was optimal at pH 4.5. The Km value with galactoglucomannan as substrate was 0.50 mg ml?1. One xylanase with molecular mass of 79 kDa was also purified and characterized. Its activity was optimal at 60 °C and pH 8.0. Its Km value with birchwood xylan as substrate was 1.65 mg ml?1. Both the mannanase and the 79 kDa xylanase displayed relatively high activity on carboxymethyl cellulose. The sensitivity of the xylanase and mannanase to various salts was evaluated. None of the tested salts inhibited the xylanase, but Mn+2, Fe+3, and Cu+2 were strong inhibitors for the mannanase.  相似文献   

2.
The present study was undertaken to gain insight into the associations of mercury(II) with dicysteinyl tripeptides in buffered media at pH 7.4. We investigated the effects of increasing the distance between cysteinyl residues on mercury(II) associations and complex formations. The peptide–mercury(II) formation constants and their associated thermodynamic parameters in 3-(N-morpholino)propanesulfonic acid (MOPS) buffered solutions were evaluated by isothermal titration calorimetry. Complexes formed in different relative ratios of mercury(II) to cysteinyl peptides in ammonium formate buffered solutions were characterized by LTQ Orbitrap mass spectrometry. The results from these studies show that n-alkyl dicysteinyl peptides (CP 14), and an aryl dicysteinyl peptide (CP 5) can serve as effective “double anchors” to accommodate the coordination sites of mercury(II) to form predominantly one-to-one Hg(peptide) complexes. The aryl dicysteinyl peptide (CP 5) also forms the two-to-two Hg2(peptide)2 complex. In the presence of excess peptide, Hg(peptide)2 complexes are also detected. Notably, increasing the distance between the ligating groups or “anchor points” in CP 15 does not significantly affect their affinity for mercury(II). However, the enthalpy change (ΔH) values (ΔH1  −91 kJ mol−1 and ΔH2  −66 kJ mol−1) for complex formation between CP 4 and 5 with mercury(II) are about one and a half times larger than the related values for CP 1, 2 and 3H1  −66 kJ mol−1 and ΔH2  46 kJ mol−1). The corresponding entropy change (ΔS) values (ΔS1  −129 J K−1 mol−1 and ΔS2  −116 J K−1 mol−1) of the structurally larger dicysteinyl peptides CP 4 and 5 are less entropically favorable than for CP 1, 2 and 3S1  −48 J K−1 mol−1 and ΔS2  −44 J K−1 mol−1). Generally, these associations result in a decrease in entropy, indicating that these peptide–mercury complexes potentially form highly ordered structures. The results from this study show that dicysteinyl tripeptides are effective in binding mercury(II) and they are promising motifs for the design of multi-cysteinyl peptides for binding more than one mercury(II) ion per peptide.  相似文献   

3.
The studies on adsorption of hexavalent chromium were conducted by varying various parameters such as contact time, pH, amount of adsorbent, concentration of adsorbate and temperature. The kinetics of adsorption of Cr(VI) ion followed pseudo second order. Langmuir adsorption isotherm was employed in order to evaluate the optimum adsorption capacity of the adsorbent. The adsorption capacity was found to be pH dependant. Sawdust was found to be very effective and reached equilibrium in 3 h (adsorbate concentration 30 mg l−1). The rate constant has been calculated at 303, 308, 313 and 318 K and the activation energy (Ea) was calculated using the Arrhenius equation. Thermodynamic parameters such as standard Gibbs energy (ΔG°) and heat of adsorption (ΔHr) were calculated. The ΔG° and ΔHr values for Cr(VI) adsorption on the sawdust showed the process to be exothermic in nature. The percentage of adsorption increased with decrease in pH and showed maximum removal of Cr(VI) in the pH range 4.5–6.5 for an initial concentration of 5 mg l−1.  相似文献   

4.
β-Glucosidase catalyzes the sequential breakdown of cyanogenic glycosides in cyanogenic plants. The β-glucosidase from Prunus armeniaca L. was purified to 8-fold, and 20% yield was obtained, with a specific activity of 281 U/mg protein. The enzyme showed maximum activity in 0.15 M sodium citrate buffer, pH 6, at 35 °C with p-nitrophenylglucopyranoside as substrate. The β-glucosidase from wild apricot was used successfully for the saccharification of cellobiose into D-glucose. This enzyme has a Vmax of 131.6 μmol min−1 mg−1 protein, Km of 0.158 mM, Kcat of 144.8 s−1, Kcat/Km of 917.4 mM−1 s−1, and Km/Vmax of 0.0012 mM min mg μmole−1, using cellobiose as substrate. The half-life, deactivation rate coefficient, and activation energy of this β-glucosidase were 12.76 h, 1.509 × 10−5 s−1, and 37.55 kJ/mol, respectively. These results showed that P. armeniaca is a potential source of β-glucosidase, with high affinity and catalytic capability for the saccharification of cellulosic material.  相似文献   

5.
In the current work nanoparticles (NPs) of α-amylase were generated in an aqueous solution using high-intensity ultrasound, and were subsequently immobilized on polyethylene (PE) films, or polycarbonate (PC) plates, or on microscope glass slides. The α-amylase NPs coated on the solid surfaces have been characterized by ESEM, TEM, FTIR, XPS and AFM. The substrates immobilized with α-amylase were used for hydrolyzing soluble potato starch to maltose. The amount of enzyme introduced in the substrates, leaching properties, and the catalytic activity of the immobilized enzyme were compared. The catalytic activity of the amylase deposited on the three solid surfaces was compared to that of the same amount of free enzyme at different pHs and temperatures. α-Amylase coated on PE showed the best catalytic activity in all the examined parameters when compared to native amylase, especially at high temperatures. When immobilized on glass, α-amylase showed better activity than the native enzyme over all pH and temperature values studied. However, the immobilization on PC did not improve the enzyme activity at any pH and any temperature compared to the free amylase. The kinetic parameters, Km and Vmax were also calculated. The amylase coated PE showed the most favorable kinetic parameters (Km = 5 g L−1 and Vmax = 5E−07 mol mL−1 min−1). In contrast, the anchored enzyme-PC exhibited unfavorable kinetic parameters (Km = 16 g L−1, Vmax = 4.2E−07 mol mL−1 min−1). The corresponding values for amylase-glass were Km = 7 g L−1, Vmax = 1.8E−07 mol mL−1 min−1, relative to those obtained for the free enzyme (Km = 6.6 g L−1, Vmax = 3.3E−07 mol mL−1 min−1).  相似文献   

6.
Aerobic granulation is a process in which suspended biomass aggregate and form discrete well-defined granules in aerobic systems. To investigate the properties and kinetics of aerobic granular sludge, aerobic granules were cultivated with glucose synthetic wastewater in a series of sequencing batch reactors (SBR). The spherical shaped granules were observed on 8th day with the mean diameter of 0.1 mm. With the organic loading rate (OLR) being increased to 4.0 g COD L−1 d−1, aerobic granules grew matured with spherical shape. The size of granules ranged from 1.2 to 1.8 mm, and the corresponding settling velocity of individual granule was 24.2–36.4 m h−1. The oxygen utilization rate (OUR) of mature granules was 41.90 g O2 kg MLSS−1 h−1, which was two times higher than that of activated sludge (18.32 g O2 kg MLSS−1 h−1). The experimental data indicated that the substrate utilization and biomass growth kinetics generally followed Monod's kinetics model. The corresponding kinetic coefficients of k (maximum specific substrate utilization rate), Ks (half velocity coefficient), Y (growth yield coefficient) and Kd (decay coefficient) were determined as follows, kc = 23.65 d−1, Kc = 3367.05 mg L−1, KN = 0.038 d−1, KN = 29.65 mg L−1, Y = 0.1927–0.2022 mg MMLS (mg COD)−1 and Kd = 0.00845–0.0135 d−1, respectively. Those properties of aerobic granules made aerobic granules system had a short setup period, high substrate utilization rate and low sludge production.  相似文献   

7.
Glycine oxidase (GO) has great potential for use in biosensors, industrial catalysis and agricultural biotechnology. In this study, a novel GO (BliGO) from a marine bacteria Bacillus licheniformis was cloned and characterized. BliGO showed 62% similarity to the well-studied GO from Bacillus subtilis. The optimal activity of BliGO was observed at pH 8.5 and 40 °C. Interestingly, BliGO retained 60% of the maximum activity at 0 °C, suggesting it is a cold-adapted enzyme. The kinetic parameters on glyphosate (Km, kcat and kcat/Km) of BliGO were 11.22 mM, 0.08 s−1, and 0.01 mM−1 s−1, respectively. To improve the catalytic activity to glyphosate, the BliGO was engineered by directed evolution. With error-prone PCR and two rounds of DNA shuffling, the most evolved mutant SCF-4 was obtained from 45,000 colonies, which showed 7.1- and 8-fold increase of affinity (1.58 mM) and catalytic efficiency (0.08 mM−1 s−1) to glyphosate, respectively. In contrast, its activity to glycine (the natural substrate of GO) decreased by 113-fold. Structure modeling and site-directed mutation study indicated that Ser51 in SCF-4 involved in the binding of enzyme with glyphosate and played a crucial role in the improvement of catalytic efficiency.  相似文献   

8.
The introduced shrub Tamarix ramosissima invades riparian zones, but loses competitiveness under flooding. Metabolic effects of flooding could be important for T. ramosissima, but have not been previously investigated. Photosynthesis rates, stomatal conductance, internal (intercellular) CO2, transpiration, and root alcohol dehydrogenase (ADH) activity were compared in T. ramosissima across soil types and under drained and flooded conditions in a greenhouse. Photosynthesis at 1500 μmol quanta m−2 s−1 (A1500) in flooded plants ranged from 2.3 to 6.2 μmol CO2 m−2 s−1 during the first week, but A1500 increased to 6.4–12.7 μmol CO2 m−2 s−1 by the third week of flooding. Stomatal conductance (gs) at 1500 μmol quanta m−2 s−1 also decreased initially during flooding, where gs was 0.018 to 0.099 mol H2O m−2 s−1 during the first week, but gs increased to 0.113–0.248 mol H2O m−2 s−1 by the third week of flooding. However, photosynthesis in flooded plants was reduced by non-stomatal limitations, and subsequent increases indicate metabolic acclimation to flooding. Root ADH activities were higher in flooded plants compared to drained plants, indicating oxygen stress. Lower photosynthesis and greater oxygen stress could account for the susceptibility of T. ramosissima at the onset of flooding. Soil type had no effect on photosynthesis or on root ADH activity. In the field, stomatal conductance, leaf water potential, transpiration, and leaf δ13C were compared between T. ramosissima and other flooded species. T. ramosissima had lower stomatal conductance and water potential compared to Populus deltoides and Phragmites australis. Differences in physiological responses for T. ramosissima could become important for ecological concerns.  相似文献   

9.
Apurinic/apyrimidinic (AP) endonucleases are important DNA repair enzymes involved in two overlapping pathways: DNA glycosylase-initiated base excision (BER) and AP endonuclease-initiated nucleotide incision repair (NIR). In the BER pathway, AP endonucleases cleave DNA at AP sites and 3'-blocking moieties generated by DNA glycosylases, whereas in NIR, the same AP endonucleases incise DNA 5' to a wide variety of oxidized bases. The flowering plant Arabidopsis thaliana contains three genes encoding homologues of major human AP endonuclease 1 (APE1): Arp, Ape1L and Ape2. It has been shown that all three proteins contain AP site cleavage and 3'-repair phosphodiesterase activities; however, it was not known whether the plant AP endonucleases contain the NIR activity. Here, we report that ARP proteins from Arabidopsis and common wheat (Triticum aestivum) contain NIR and 3'  5' exonuclease activities in addition to their AP endonuclease and 3'-repair phosphodiesterase functions. The steady-state kinetic parameters of reactions indicate that Arabidopsis ARP cleaves oligonucleotide duplexes containing α-anomeric 2'-deoxyadenosine (αdA) and 5,6-dihydrouridine (DHU) with efficiencies (kcat/KM = 134 and 7.3 μM−1·min−1, respectively) comparable to those of the human counterpart. However, the ARP-catalyzed 3'-repair phosphodiesterase and 3'  5' exonuclease activities (kcat/KM = 314 and 34 μM−1·min−1, respectively) were about 10-fold less efficient as compared to those of APE1. Interestingly, homozygous A. thaliana arp–/– mutant exhibited high sensitivity to methyl methanesulfonate and tert-butyl hydroperoxide, but not to H2O2, suggesting that ARP is a major plant AP endonuclease that removes abasic sites and specific types of oxidative DNA base damage. Taken together, these data establish the presence of the NIR pathway in plants and suggest its possible role in the repair of DNA damage generated by oxidative stress.  相似文献   

10.
Marine toxic dinoflagellates of the genus Gambierdiscus are the causative agents of ciguatera fish poisoning (CFP), a form of seafood poisoning that is widespread in tropical, subtropical and temperate regions worldwide. The distributions of Gambierdiscus australes, Gambierdiscus scabrosus and two phylotypes of Gambierdiscus spp. type 2 and type 3 have been reported for the waters surrounding the main island of Japan. To explore the bloom dynamics and the vertical distribution of these Japanese species and phylotypes of Gambierdiscus, the effects of light intensity on their growth were tested, using a photoirradiation-culture system. The relationship between the observed growth rates and light intensity conditions for the four species/phylotypes were formulated at R > 0.92 (p < 0.01) using regression analysis and photosynthesis-light intensity (P-L) model. Based on this equation, the optimum light intensity (Lmax) and the semi-optimum light intensity range (Ls-opt) that resulted in the maximum growth rate (μmax) and ≥80% μ max values of the four species/phylotypes, respectively, were as follows: (1) the Lmax and Ls-opt of G. australes were 208 μmol photons m−2 s−1 and 91–422 μmol photons m−2 s−1, respectively; (2) those of G. scabrosus were 252 and 120–421 μmol photons m−2 s−1, respectively; (3) those of Gambierdiscus sp. type 2 were 192 and 75–430 μmol photons m−2 s−1, respectively; and (4) those of Gambierdiscus sp. type 3 were ≥427 and 73–427 μmol photons m−2 s−1, respectively. All four Gambierdiscus species/phylotypes required approximately 10 μmol photons m−2 s−1 to maintain growth. The light intensities in coastal waters at a site in Tosa Bay were measured vertically at 1 m intervals once per season. The relationships between the observed light intensity and depth were formulated using Beer’s Law. Based on these equations, the range of the attenuation coefficients at Tosa Bay site was determined to be 0.058–0.119 m−1. The values 1700 μmol photons m−2 s−1, 500 μmol photons m−2 s−1, and 200 μmol photons m−2 s−1 were substituted into the equations to estimate the vertical profiles of light intensity at sunny midday, cloudy midday and rainy midday, respectively. Based on the regression equations coupled with the empirically determined attenuation coefficients for each of the four seasons, the ranges of the projected depths of Lmax and Ls-opt for the four Gambierdiscus species/phylotypes under sunny midday conditions, cloudy midday conditions, and rainy midday conditions were 12–38 m and 12–54 m, 1–16 m and 1–33 m, and 0 m and 0–16 m, respectively. These results suggest that light intensity plays an important role in the bloom dynamics and vertical distribution of Gambierdiscus species/phylotypes in Japanese coastal waters.  相似文献   

11.
《Process Biochemistry》2010,45(1):88-93
A fibrinolytic protease (FP84) was purified from Streptomyces sp. CS684, with the aim of isolating economically viable enzyme from a microbial source. SDS-PAGE and fibrin zymography of the purified enzyme showed a single protein band of approximately 35 kDa. Maximal activity was at 45 °C and pH 7–8, and the enzyme was stable between pH 6 and 9 and below 40 °C. It exhibited fibrinolytic activity, which is stronger than that of plasmin. FP84 hydrolyzed Bβ-chains of fibrinogen, but did not cleave Aα- and γ-chains. Km, Vmax and Kcat values for azocasein were 4.2 mg ml−1, 305.8 μg min−1 mg−1 and 188.7 s−1, respectively. The activity was suppressed by Co2+, Zn2+, Cu2+ and Fe2+, but slightly enhanced by Ca2+ and Mg+2. Additionally, the activity was slightly inhibited by aprotinin and PMSF, but significantly inhibited by pefabloc, EDTA and EGTA. The first 15 amino acids of N-terminal sequence were GTQENPPSSGLDDID. They are highly similar to those of serine proteases from various Streptomyces strains, but different with known fibrinolytic enzymes. These results suggest that FP84 is a novel serine metalloprotease with potential application in thrombolytic therapy.  相似文献   

12.
《Process Biochemistry》2010,45(7):1036-1042
A recombinant strain of Escherichia coli with CYP102A1 gene was developed for the demethylation of colchicine into their derivatives. The CYP102A1 gene responsible for demethylation was isolated from Bacillus megaterium ACBT03 and amplified using suitable primers. The amplified product was cloned into pET28a+ expression vector using host E. coli BL21(DE3) cells. The CYP3A4 (product of CYP102A1 gene) protein expression and other parameters like substrate toxicity, product toxicity and enzyme activity were optimized in shake flasks; and further scaled-up to 5 l bioreactor with 3 l working volume. In 5 l bioreactor, dissolved oxygen (DO) was optimized for maximum specific growth and enhanced 3-demethylated colchicine (3-DMC) production. The optimized conditions from shake flasks were scaled-up to 70 l bioreactor and resulted into ∼80% conversion of 20 mM colchicine in 48 h with a volumetric productivity of 6.62 mg l−1 h−1. Scale-up factors were measured as volumetric oxygen transfer coefficient (kLa) i.e., 56 h−1 and impeller tip velocity (Vtip) i.e., 7.065 m s−1, respectively. The kinetic parameters Km, kcat, and kcat/Km of the CYP3A4 enzyme using colchicine as the substrate were determined to be 271 ± 30 μM, 8533 ± 25 min−1, and 31.49 μM min−1, respectively, when IPTG induced recombinant E. coli culture was used.  相似文献   

13.
Leifsonia xyli HS0904 can stereoselectively catalyze the bioreduction of 3,5-bis(trifluoromethyl) acetophenone (BTAP) to its corresponding alcohol, which is a valuable chiral intermediate in the pharmaceuticals. In this study, a new carbonyl reductase derived from L. xyli HS0904 was purified and its biochemical properties were determined in detail. The carbonyl reductase was purified by 530-fold with a specific activity of 13.2 U mg−1 and found to be a homodimer with a molecular mass of 49 kDa, in which the subunit molecular-weight was about 24 kDa. The purified enzyme exhibited a maximum enzyme activity at 34 °C and pH 7.2, and retained over 90% of its initial activity at 4 °C and pH 7.0 for 24 h. The addition of various additives, such as Ca2+, Mg2+, Mn2+, l-cysteine, l-glutathione, urea, PEG 1000 and PEG 4000, could enhance the enzyme activity. The maximal reaction rate (Vmax) and apparent Michaelis–Menten constant (Km) of the purified carbonyl reductase for BTAP and NADH were confirmed as 33.9 U mg−1, 0.383 mM and 69.9 U mg−1, 0.412 mM, respectively. Furthermore, this enzyme was found to have a broad spectrum of substrate specificity and can asymmetrically catalyze the reduction of a variety of ketones and keto esters.  相似文献   

14.
《Aquatic Botany》2002,72(3-4):219-233
We studied the potential role of dissolved inorganic carbon (DIC) in determining vegetation dominance of Potamogeton pectinatus L. and Chara aspera Deth. ex Willd. by monitoring the seasonal dynamics of DIC in a shallow lake and comparing the use of DIC of the two species. The HCO3-concentration in summer dropped from 2.5 to <0.5 mM with seasonally increasing Chara biomass, whereas outside the vegetation concentrations remained at 2.5 mM. Inside Potamogeton spp. vegetation DIC decreased from 2.5 to ca. 0.75 mM HCO3. A growth experiment showed ash-free biomass for P. pectinatus was nearly two times as high as for C. aspera at 3 mM HCO3, but almost two times lower at 0.5 mM than at 3.0. In a separate experiment, P. pectinatus precultured at a relatively low HCO3-level had a lower net photosynthetic rate (Pmax, 0.1 mmol O2 g−1 DW h−1) than C. aspera (Pmax, 0.1 mmol O2 g−1 DW h−1) over the range of HCO3-concentrations tested (Pmax, 0.14 mmol O2 g−1 DW h−1). In response to CO2 no significant differences between the compensation points (P. pectinatus, 28 mM; C. aspera 66 mM), were observed, but the photosynthetic rate increased faster than for C. aspera than for P. pectinatus. Under field conditions, the use of CO2 is not important since inside vegetation CO2-concentrations were below 10 μM, and thus, not available for photosynthesis of either species during the main part of the growth season. It is suggested that C. aspera may be a better competitor for HCO3 than P. pectinatus in conditions with a low HCO3 supply. As HCO3 is a strong limiting factor for growth inside the vegetation and probably the only carbon source available, the superior ability of C. aspera to use HCO3 may be an important factor explaining its present dominance in Veluwemeer.  相似文献   

15.
Salts inhibit the activity of sweet almond β-glucosidase. For cations (Cl salts) the effectiveness follows the series: Cu+2, Fe+2 > Zn+2 > Li+ > Ca+2 > Mg+2 > Cs+ > NH4+ > Rb+ > K+ > Na+ and for anions (Na+ salts) the series is: I > ClO4 > SCN > Br  NO3 > Cl  OAc > F  SO4 2. The activity of the enzyme, like that of most glycohydrolases, depends on a deprotonated carboxylate (nucleophile) and a protonated carboxylic acid for optimal activity. The resulting pH-profile of kcat/Km for the β-glucosidase-catalyzed hydrolysis of p-nitrophenyl glucoside is characterized by a width at half height that is strongly sensitive to the nature and concentration of the salt. Most of the inhibition is due to a shift in the enzymic pKas and not to an effect on the pH-independent second-order rate constant, (kcat/Km)lim. For example, as the NaCl concentration is increased from 0.01 M to 1.0 M the apparent pKa1 increases (from 3.7 to 4.9) and the apparent pKa2 decreases (from 7.2 to 5.9). With p-nitrophenyl glucoside, the value of the pH-independent (kcat/Km)lim (= 9 × 104 M 1 s 1) is reduced by less than 4% as the NaCl concentration is increased. There is a similar shift in the pKas when the LiCl concentration is increased to 1.0 M. The results of these salt-induced pKa shifts rule out a significant contribution of reverse protonation to the catalytic efficiency of the enzyme. At low salt concentration, the fraction of the catalytically active monoprotonated enzyme in the reverse protonated form (i.e., proton on the group with a pKa of 3.7 and dissociated from the group with a pKa of 7.2) is very small (≈ 0.03%). At higher salt concentrations, where the two pKas become closer, the fraction of the monoprotonated enzyme in the reverse protonated form increases over 300-fold. However, there is no increase in the intrinsic reactivity, (kcat/Km)lim, of the monoprotonated species. For other enzymes which may show such salt-induced pKa shifts, this provides a convenient test for the role of reverse protonation.  相似文献   

16.
We have cloned, purified and investigated the catalytic activity and anion inhibition profiles of a full catalytic domain (358 amino acid residues) carbonic anhydrase (CA, EC 4.2.1.1) from Plasmodium falciparum, PfCAdom, an enzyme belonging to the η-CA class and identified in the genome of the malaria-producing protozoa. A truncated such enzyme, PfCA1, containing 235 residues was investigated earlier for its catalytic and inhibition profiles. The two enzymes were efficient catalysts for CO2 hydration: PfCAdom showed a kcat of 3.8 × 105 s−1 and kcat/Km of 7.2 × 107 M−1 × s−1, whereas PfCA showed a lower activity compared to PfCAdom, with a kcat of 1.4 × 105 s−1 and kcat/Km of 5.4 × 106 M−1 × s−1. PfCAdom was generally less inhibited by most anions and small molecules compared to PfCA1. The best PfCAdom inhibitors were sulfamide, sulfamic acid, phenylboronic acid and phenylarsonic acid, which showed KIs in the range of 9–68 μM, followed by bicarbonate, hydrogensulfide, stannate and N,N-diethyldithiocarbamate, which were submillimolar inhibitors, with KIs in the range of 0.53–0.97 mM. Malaria parasites CA inhibition was proposed as a new strategy to develop antimalarial drugs, with a novel mechanism of action.  相似文献   

17.
The biogeochemical cycles of nitrogen (N) and base cations (BCs), (i.e., K+, Na+, Ca2+, and Mg2+), play critical roles in plant nutrition and ecosystem function. Empirical correlations between large experimental N fertilizer additions to forest ecosystems and increased BCs loss in stream water are well demonstrated, but the mechanisms driving this coupling remain poorly understood. We hypothesized that protons generated through N transformation (PPRN)—quantified as the balance of NH4+ (H+ source) and NO3 (H+ sink) in precipitation versus the stream output will impact BCs loss in acid-sensitive ecosystems. To test this hypothesis, we monitored precipitation input and stream export of inorganic N and BCs for three years in an acid-sensitive forested watershed in a granite area of subtropical China. We found the precipitation input of inorganic N (17.71 kg N ha−1 year−1 with 54% as NH4+–N) was considerably higher than stream exported inorganic N (5.99 kg N ha−1 year−1 with 83% as NO3–N), making the watershed a net N sink. The stream export of BCs (151, 1518, 851, and 252 mol ha−1 year−1 for K+, Na+, Ca2+, and Mg2+, respectively) was positively correlated (r = 0.80, 0.90, 0.84, and 0.84 for K+, Na+, Ca2+, and Mg2+ on a monthly scale, respectively, P < 0.001, n = 36) with PPRN (389 mol ha−1 year−1) over the three years, suggesting that PPRN drives loss of BCs in the acid-sensitive ecosystem. A global meta-analysis of 15 watershed studies from non-calcareous ecosystems further supports this hypothesis by showing a similarly strong correlation between ∑BCs output and PPRN (r = 0.89, P < 0.001, n = 15), in spite of the pronounced differences in environmental settings. Collectively, our results suggest that N transformations rather than anions (NO3 and/or SO42−) leaching specifically, are an important mediator of BCs loss in acid-senstive ecosystems. Our study provides the first definitive evidence that the chronic N deposition and subsequent transformation within the watershed drive stream export of BCs through proton production in acid-sensitive ecosystems, irrespective of their current relatively high N retention. Our findings suggest the N-transformation-based proton production can be used as an indicator of watershed outflow quality in the acid-sensitive ecosystems.  相似文献   

18.
《Aquatic Botany》2005,81(2):157-173
The main photosynthesis and respiration parameters (dark respiration rate, light saturated production rate, saturation irradiance, photosynthetic efficiency) were measured on a total of 23 macrophytes of the Thau lagoon (2 Phanerogams, 5 Chlorophyceae, 10 Rhodophyceae and 6 Phaeophyceae). Those measurements were performed in vitro under controlled conditions, close to the natural ones, and at several seasons. Concomitantly, measurements of pigment concentrations, carbon, phosphorous and nitrogen contents in tissues were performed. Seasonal intra-specific variability of photosynthetic parameters was found very high, enlightening an important acclimatation capacity. The highest photosynthetic capacities were found for Chlorophyceae (e.g. Monostroma obscurum thalli at 17 °C, 982 μmol O2 g−1 dw h−1 and 9.1 μmol O2 g−1 dw h−1/μmol photons m−2 s−1, respectively for light saturated net production rate and photosynthetic efficiency) and Phanerogams (e.g. Nanozostera noltii leaves at 25 °C, 583 μmol O2 g−1 dw h−1 and 2.6 μmol O2 g−1 dw h−1/μmol photons m−2 s−1 respectively for light saturated net production rate and photosynthetic efficiency). As expected, species with a high surface/volume ratio were found to be more productive than coarsely branched thalli and thick blades shaped species. Contrary to Rd (ranging 6.7–794 μmol O2 g−1 dw h−1, respectively for Rytiphlaea tinctoria at 7 °C and for Dasya sessilis at 25 °C) for which a positive relationship with water temperature was found whatever the species studied, the evolution of P/I curves with temperature exhibited different responses amongst the species. The results allowed to show summer nitrogen limitation for some species (Gracilaria bursa-pastoris and Ulva spp.) and to propose temperature preferences based on the photosynthetic parameters for some others (N. noltii, Zostera marina, Chaetomorpha linum).  相似文献   

19.
The inhibitory effect of ammonium sulfate on a commercial mixed culture, used in biological waste-water treatment was studied under aerobic batch conditions. Several mathematical models of enzyme and growth kinetics including a death factor were analyzed through nonlinear regression to find the best fit to corresponding data of inhibition. The best fit model was found to be the generalized Monod type with a death factor having the biokinetic parameters; μmax 0.681 h−1, Ks 0.224 g dm−3, Ki 56240 g dm−3, K 0.055 g dm−3 and kd 0.052 h−1 to represent the experimental data accurately. The low saturation coefficient value along with high maximum specific growth rate and inhibition coefficient denotes the competitive characteristics of commercial mixed cultures in the biological treatment of high ammonium polluted waste waters.  相似文献   

20.
《Aquatic Botany》2005,81(4):326-342
The effects of NH4+ or NO3 on growth, resource allocation and nitrogen (N) uptake kinetics of two common helophytes Phragmites australis (Cav.) Trin. ex Steudel and Glyceria maxima (Hartm.) Holmb. were studied in semi steady-state hydroponic cultures. At a steady-state nitrogen availability of 34 μM the growth rate of Phragmites was not affected by the N form (mean RGR = 35.4 mg g−1 d−1), whereas the growth rate of Glyceria was 16% higher in NH4+-N cultures than in NO3-N cultures (mean = 66.7 and 57.4 mg g−1 d−1 of NH4+ and NO3 treated plants, respectively). Phragmites and Glyceria had higher S/R ratio in NH4+ cultures than in NO3 cultures, 123.5 and 129.7%, respectively.Species differed in the nitrogen utilisation. In Glyceria, the relative tissue N content was higher than in Phragmites and was increased in NH4+ treated plants by 16%. The tissue NH4+ concentration (mean = 1.6 μmol g fresh wt−1) was not affected by N treatment, whereas NO3 contents were higher in NO3 (mean = 1.5 μmol g fresh wt−1) than in NH4+ (mean = 0.4 μmol g fresh wt−1) treated plants. In Phragmites, NH4+ (mean = 1.6 μmol g fresh wt−1) and NO3 (mean = 0.2 μmol g fresh wt−1) contents were not affected by the N regime. Species did not differ in NH4+ (mean = 56.5 μmol g−1 root dry wt h−1) and NO3 (mean = 34.5 μmol g−1 root dry wt h−1) maximum uptake rates (Vmax), and Vmax for NH4+ uptake was not affected by N treatment. The uptake rate of NO3 was low in NH4+ treated plants, and an induction phase for NO3 was observed in NH4+ treated Phragmites but not in Glyceria. Phragmites had low Km (mean = 4.5 μM) and high affinity (10.3 l g−1 root dry wt h−1) for both ions compared to Glyceria (Km = 6.3 μM, affinity = 8.0 l g−1 root dry wt h−1). The results showed different plasticity of Phragmites and Glyceria toward N source. The positive response to NH4+-N source may participates in the observed success of Glyceria at NH4+ rich sites, although other factors have to be considered. Higher plasticity of Phragmites toward low nutrient availability may favour this species at oligotrophic sites.  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