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.
In order to better understand the relative importance of different ecosystems and nitrogen cycling processes within the Amazon basin to the nitrogen economy of this region, we constructed a generalized nitrogen budget for the region based on data for hydrologic losses of nitrogen and nitrogen fixation in Amazon forests. Data included information available for nitrogen in water entering and leaving both the entire basin and watersheds on oxisol and ultisol soils near Manaus, Brazil, in addition to biological nitrogen fixation in forests on ultisol, oxisol and entisol (‘varzea’) soils in Central Amazonia. Available data indicate that 4–6 kg N ha?1 yr?1 are lost via the River Amazonas, and that a similar amount enters in rainfall. Root-associated biological nitrogen fixation contributesca. 2 kg N ha?1 yr?1 to forests on oxisols, 20 kg N ha?1 yr?1 to forests on utisols, and 200 kg N ha?1 yr?1 to forests on fertile varzea soils. There is 5–10 fold more NH4 +?N than NO3?N in rain and stream water entering and leaving the waterbasin near Manaus. Calculations based on these data plus certain assumption yield the following regional nitrogen balance estimate: inputs through bulk deposition of 36×108 kg N yr?1 and through biological nitrogen fixation of 120×108 kg N yr?1, and outputsvia the River Amazonas of 36×108 kg N yr?1 andvia denitrification and volatization (by difference) of 120×108 kg N yr?1.  相似文献   

2.
Lake Muzahi,Rwanda: limnological features and phytoplankton production   总被引:1,自引:1,他引:0  
Lake Muhazi, a small lake of Rwanda (East Africa) was studied from 1986 to 1990. A dramatic decrease of the catch of Oreochromis niloticus (350 T y−1 in the fifties vs 30 T y−1 in 1982) suggested a loss of productivity or overfishing. In the same period, other ecological changes occurred: the submerged macrophytes regressed and there was a decrease in Secchi depth (0.65 m in 1987 vs 1.5 m in the fifties). Compared to other lakes of the same area, the plankton production seemed low. The results of the present study characterize lake Muhazi as a shallow lake with a rather unstable diurnal stratification and with slight differences in mixing regime between its eastern, deepest part and its western, shallowest part. Secchi disk depth does not vary seasonally to a large extent. The water has a rather high mineral content (conductivity of about 500 μS cm−1 at 25 °C) and low concentrations of dissolved N and P, except in the hypolimnion, where NH inf4 sup+ -N can be high. Two species, Microcystis aeruginosa and Ceratium hirundinella, account for most of the phytoplankton biomass, which is about 50–80 mg chlorophyll a m−2 in the euphotic zone, usually with little seasonal variation. Daily gross production estimates amount to about 6 to 9.5 g O2 m−2 d−1 with a significant difference between the two parts of the lake. Data on C:N and C:P ratio in the phytoplankton suggest that some N deficiency might occur in the eastern part. Moreover, the Zm:Zc ratio could also lead to rather low net production rates (0.21–0.25 d−1 for a mixed layer of 4 m) In conclusion, the primary production of lake Muhazi is medium for African lakes and the hypothesis that decreased planktonic production could account for a reduced fish production should be discarded. Whereas the present yield of the fishery is only 20 kg ha−1 y−1, the yield estimated from primary production ranges between 46 and 64 kg ha−1 y−1. This could be reached through proper management. Finally, some hypotheses are given to explain the ecological changes which occurred in the lake.  相似文献   

3.
Diatom cell quantity and their biochemical composition vary among species and are greatly affected by harvest stage or culture conditions. This study compares growth pattern, cell attachment, and biochemical composition of four diatoms suitable for abalone post-larvae: Navicula incerta, Proschkinia sp., Nitzschia sp., and Amphora sp. The four diatoms were grown in F/2 medium at 28.5?±?1.4°C, under 62?±?8?μmol?photons?m?2?s?1, at different original inoculating densities (0.05?×?106, 0.10?×?106, and 0.25?×?106?cells?mL?1) and were harvested in log and stationary phase of growth for biochemical analysis. Total protein, carbohydrate, lipid, and ash composition, as well as fatty acid composition, were determined. All diatoms grew better when inoculated at 0.10?×?106?cell?mL?1 with Proschkinia sp. reaching the highest cell density of 6.56?×?106?cells?mL?1 in log phase. Amphora sp. had the highest cell attachment capacity when inoculated at 0.10?×?106?cell?mL?1 (11,580?cells?mm?2), whereas N. incerta had the lowest (7,750?cells?mm?2). Protein and lipid (percent dry weight) contents were generally highest in cells during log phase of growth; Amphora sp. in log phase of growth had the highest lipid content of 9.74% DW, whereas significant differences in carbohydrate between the two growth phases were only observed for Proschkinia sp. Besides, all diatoms had higher energy contents in log phase of growth. There were no significant differences in ash content among the four diatoms. Polyunsaturated fatty acid (PUFA) content ranged from 23.25% to 38.62% of the total fatty acids, and the four diatoms tested were richer in n-3 PUFA than in n-6 PUFA. All the diatoms had significant quantities of 20:5n-3 (EPA) (between 12.69% and 17.68% of TFA), and Proschkinia sp., in log phase of growth, had the highest quantity of arachidonic acid (20:4n-6; ARA). The results highlight the influence of culture conditions and harvest protocols on diatom nutritive value and enabled a preliminary approach towards the selection of novel diatom species.  相似文献   

4.
Fine roots <2 mm in diameter play a key role in regulating the biogeochemical cycles of ecosystems and are important to our understanding of ecosystem responses to global climate changes. Given the sensitivity of fine roots, especially in boreal region, to climate changes, it is important to assess whether and to what extent fine roots in this region change with climates. Here, in this synthesis, a data set of 218 root studies were complied to examine fine root patterns in the boreal forest in relation to site and climatic factors. The mean fine root biomass in the boreal forest was 5.28 Mg ha?1, and the production of fine roots was 2.82 Mg ha?1 yr?1, accounting for 32% of annual net primary production of the boreal forest. Fine roots in the boreal forest on average turned over 1.07 times per year. Fine roots contained 50.9 kg ha?1 of nitrogen (N) and 3.63 kg ha?1 of phosphorous (P). In total, fine roots in the boreal forest ecosystems contain 6.1 × 107 Mg N and 4.4×106Mg P pools, respectively, about 10% of the global nutrients of fine roots. Fine root biomass, production, and turnover rate generally increased with increasing mean annual temperature and precipitation. Fine root biomass in the boreal forest decreased significantly with soil N and P availability. With increasing stand age, fine root biomass increased until about 100 years old for forest stands and then leveled off or decreased thereafter. These results of meta analysis suggest that environmental factors strongly influence fine root biomass, production, and turnover in boreal forest, and future studies should place a particular emphasis on the root-environment relationships.  相似文献   

5.
1. Microbial plankton dynamics in an ultra‐oligotrophic epishelf lake (Beaver Lake, Antarctica) were investigated over an austral summer (December 2002 to January 2003). The aim was to characterise carbon cycling in an environmentally extreme lake. 2. The lake had an unusual temperature profile with peak temperatures of 1.3–1.9 °C between 20 and 25 m. Photosynthetically active radiation penetrated to the lake bottom (110 m) on occasions. The ice cover underwent marked thinning and melting during the study period. 3. Chlorophyll a concentrations were consistently low, usually below 1 μg L?1, with highest concentrations close to the lake bottom, where the photosynthetic elements showed strong autofluorescence. Mean photosynthetic nanoflagellates ranged between 34.9 × 104 L?1 ± 33.5 (23rd December) and 130.9 × 104 L?1 ± 112.3 (4th December). Highest photosynthetic activity was usually recorded below 25 m. Rates of carbon fixation varied between 0.089 μg C L?1 h?1 ± 0.002 and 0.579 μg C L?1 h?1 ± 0.156. Primary production was limited by low temperature and orthophosphate availability. 4. Mean bacterial concentration throughout the water column ranged between 9.3 × 107 L?1 ± 1.2 (23rd December) and 14.0 × 107 L?1 ± 1.8 (28th January). Bacterial production was low, less than 10% of primary production and ranged between 2.1 ng C L?1 h?1 ± 0.8 and 12 ng C L?1 h?1 ± 0.9. Highest rates coincided with times of highest primary production. On occasion dissolved organic carbon (DOC) concentrations dropped to 20 μg L?1, probably below accurate limits of detection, suggesting that carbon substratum and phosphorus may have limited bacterial growth. 5. Heterotrophic nanoflagellates varied significantly over the summer from a mean of 26.6 × 104 L?1 ± 14.2 (23rd December) to 133.8 × 104 L?1 ± 33.5 (14th December). They imposed a significant grazing impact on the bacterioplankton, removing in excess of 100% of bacterial production in December. 6. The total organic carbon pool [DOC and particulate organic carbon (POC)] was below 600 μg L?1. The ratio of DOC : POC ranged between 0.44 : 1 and 2.8 : 1 in the upper 40 m of the water column, and 1.8 : 1 and 3.7 : 1 in the lower waters. The microbial plankton contributed 1–29% of POC, thus detrital POC made up the largest fraction of the POC pool. 7. Beaver Lake is an extreme lacustrine ecosystem where heterotrophic processes occasionally appear to be carbon limited. Significant summer ice‐melt, not seen in a previous opportunistic sampling, may be having an impact on the carbon cycle.  相似文献   

6.
The C4 perennial grasses Miscanthus × giganteus and Panicum virgatum (switchgrass) are emerging bioenergy crops. They are attractive because they are productive and recycle nutrients to the overwintering belowground rhizomes, before the dry shoots are harvested. They are therefore expected to require minimal fertilizer inputs, a desirable trait for any crop. Until now, Europe has had the only long-term trials of M. × giganteus, and these have either shown no or a small effect of nitrogen fertilization, while trials of P. virgatum in the USA have shown a clear positive effect of N fertilization. This study exploited the first long-term trials of M. × giganteus in the USA, and first side-by-side comparison with P. virgatum, to test the hypothesis that N fertilization would only improve yields of the latter. A split-plot N fertility treatment (0, 67, 134, and 202 kg(N)?ha?1) was added to >5-year-old replicated stands of the two crops at seven locations on contrasting soils in the US Midwest. Averaged across all locations, M. × giganteus yields increased significantly from 23.4 Mg ha?1 with no N fertilization to 28.9 Mg ha?1 (+25 %) at a N application rate of 202 kg ha?1. P. virgatum also showed significant yield increases from 10.33 Mg ha?1 at 0 kg(N)?ha?1 to 13.6 Mg ha?1 (+32 %) at 202 kg(N)?ha?1. Both species therefore responded to N fertilization and to a similar extent. The increase per unit of added N was small compared to crops such as Zea mays and unlikely to be economically worthwhile. Nitrogen fertilization arrested most of the long-term yield decline that would otherwise have occurred in P. virgatum, but eliminated only about 40 % of the decline observed in M. × giganteus, suggesting additional causal factors for long-term yield decline in this crop. While the crops responded to nitrogen addition at some locations, they did not at others. Therefore a one-case-fits-all optimum fertilization rate cannot be prescribed.  相似文献   

7.
Hamster liver glutathione peroxidase was purified to homogeneity in three chromatographic steps and with 30% yield. The purified enzyme had a specific activity of approximately 500 μmol cumene hydroperoxide reduced/min/mg of protein at 37 °C, pH 7.6, and 0.25 mm GSH. The enzyme was shown to be a tetramer of indistinguishable subunits, the molecular weight of which was approximately 23,000 as estimated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. A single isoelectric point of 5.0 was attributed to the active enzyme. Amino acid analysis determined that selenocysteine, identified as its carboxymethyl derivative, was the only form of selenium. One residue of cysteine was found to be present in each glutathione peroxidase subunit. The presence of tryptophan was colorimetrically determined. Pseudo-first-order kinetics of inactivation of the enzyme by iodoacetate was observed at neutral pH with GSH as the only reducing agent. An optimal pH of 8.0 at 37 °C and an activation energy of 3 kcal/mol at pH 7.6 were found. A ter-uni-ping-pong mechanism was shown by the use of an integrated-rate equation. At pH 7.6, the apparent second-order rate constants for reaction of glutathione peroxidase with hydroperoxides were as follows: k1 (t-butyl hydroperoxide), 7.06 × 105 mm min?1; k1 (cumene hydroperoxide), 1.04 × 106 mm?1 min?1; k1 (p-menthane hydroperoxide), 1.2 × 106 mm?1 min?1; k1 (diisopropylbenzene hydroperoxide), 1.7 × 106 mm?1 min?1; k1 (linoleic acid hydroperoxide), 2.36 × 106 mm?1 min?1; k1 (ethyl hydroperoxide), 2.5 × 106 mm?1 min?1; and k1 (hydrogen peroxide), 2.98 × 106 mm?1 min?1. It is concluded that for bulky hydroperoxides, the more hydrophobic the substrate, the faster its reduction by glutathione peroxidase.  相似文献   

8.
The phytocystatins are inhibitors of papain-like cysteine proteinases that are implicated in defense mechanisms and the regulation of protein turnover. BCPI-1, a Brassica rapa (Chinese cabbage) phytocystatin isolated from flower buds, contains an extended C-terminal region that contains a single Cys residue at position 102. In an effort to investigate the role of the C-terminus and this Cys residue in BCPI-1 activity, purified recombinant proteins of BCPI-1, including wild-type BCPI-1 (wtBCPI-1), N-terminus BCPI-1 (BCPI-1??C), C-terminus BCPI-1 (BCPI-1??N), and BCPI-1 with a single Cys residue exchange to Ser (BCPI-1C102S), were generated and their inhibitory activities against papain were investigated. Kinetic analysis revealed that the monomeric forms of wtBCPI-1 (K i = 6.84 ± 0.3 × 10?8 M) inhibited papain more efficiently than the dimeric forms of wtBCPI-1 (K i = 1.01 ± 0.5 × 10?7 M). Experiments with recombinant BCPI-1C102S demonstrated that the dimerization of wtBCPI-1 caused by the formation of an intermolecular disulfide bond at the cysteine residue. The inhibitory activity of the recombinant proteins, except BCPI-1??N, was reduced in the pH range of 7.0?C11.5 and was highly stable over a wide range of temperatures. Thus, dimerization mediated by the cysteine residue in the extended C-terminal region and alkaline conditions reduced the inhibitory activity of BCPI-1.  相似文献   

9.
Biofuels derived from non-crop sources, such as microalgae, offer their own advantages and limitations. Despite high growth rates and lipid accumulation, microalgae cultivation still requires more energy than it produces. Furthermore, invading organisms can lower efficiency of algae production. Simple environmental changes might be able to increase algae productivity while minimizing undesired organisms like competitive algae or predatory algae grazers. Microalgae are susceptible to pH changes. In many production systems, pH is kept below 8 by CO2 addition. Here, we uncouple the effects of pH and CO2 input, by using chemical pH buffers and investigate how pH influences Nannochloropsis salina growth and lipid accumulation as well as invading organisms. We used a wide range of pH levels (5, 6, 7, 8, 9, and 10). N. salina showed highest growth rates at pH 8 and 9 (0.19?±?0.008 and 0.19?±?0.011, respectively; mean ± SD). Maximum cell densities in these treatments were reached around 21 days into the experiment (95.6?×?106?±?9?×?106 cells mL?1 for pH 8 and 92.8?×?106?±?24?×?106 cells mL?1 for pH 9). Lipid accumulation of unbuffered controls were 21.8?±?5.8 % fatty acid methyl esters content by mass, and we were unable to trigger additional significant lipid accumulation by manipulating pH levels at the beginning of stationary phase. Ciliates (grazing predators) occurred in significant higher densities at pH 6 (56.9?±?39.6?×?104 organisms mL?1) than higher pH treatments (0.1–6.8?×?104 organisms mL?1). Furthermore, the addition of buffers themselves seemed to negatively impact diatoms (algal competitors). They were more abundant in an unbuffered control (12.7?±?5.1?×?104 organisms mL?1) than any of the pH treatments (3.6–4.7?×?104 organisms mL?1). In general, pH values of 8 to 9 might be most conducive to increasing algae production and minimizing invading organisms. CO2 addition seems more valuable to algae as an inorganic carbon source and not as an essential mechanism to reduce pH.  相似文献   

10.
Stormwater detention ponds are widely utilized as control structures to manage runoff during storm events. These ponds also represent biogeochemical hotspots, where carbon (C) and nutrients can be processed and buried in sediments. This study quantified C and nutrient [nitrogen (N) and phosphorus (P)] sources and burial rates in 14 stormwater detention ponds representative of typical residential development in coastal South Carolina. Bulk sediment accumulation was directly correlated with catchment impervious surface coverage (R2 = 0.90) with sediment accumulation rates ranging from 0.06 to 0.50 cm y?1. These rates of sediment accumulation and consequent pond volume loss were lower than anticipated based on maintenance guidelines provided by the State. N-alkanes were used as biomarkers of sediment source; the derived terrestrial aquatic ratio (TARHC) index was strongly correlated with sediment accumulation rate (R2 = 0.71) which, in conjunction with high C/N ratios (16–33), suggests that terrestrial biomass drives this sediment accumulation, with relatively minimal contributions from algal derived material. This is counter to expectations that were based on the high algal productivity generally observed in stormwater ponds and previous studies of natural lakes. Sediment C and nutrient concentrations were consistent among ponds, such that differences in burial rates were a simple function of bulk sediment accumulation rate. These burial rates (C: 8.7–161 g m?2 y?1, N: 0.65–6.4 g m?2 y?1, P: 0.238–4.13 g m?2 y?1) were similar to those observed in natural lake systems, but lower than those observed in reservoirs or impoundments. Though individual ponds were small in area (930–41,000 m2), they are regionally abundant and, when mean burial rates are extrapolated to the regional scale (≈ 21,000 ponds), ultimately sequester 2.0 × 109 g C y?1, 9.5 × 107 g N y?1, and 3.7 × 107 g P y?1 in the coastal region of South Carolina alone. Stormwater ponds represent a relatively new but increasingly significant feature of the coastal landscape and, thus, are a key component in understanding how urbanization alters the transport and transformations of C and nutrients between terrestrial uplands and downstream receiving waters.  相似文献   

11.
Molecularly ordered composites of polyvinylimidazole‐[Os(bipyridine)2Cl] (PVI‐[Os(bpy)2Cl]) and glucose oxidase (GOD) are assembled inside a film of aligned carbon nanotubes. The structure of the prepared GOD/PVI‐[Os(bpy)2Cl]/CNT composite film is entirely uniform and stable; more than 90% bioelectrocatalytic activity could be maintained even after storage for 6 d. Owing to the ideal positional relationship achieved between enzyme, mediator, and electrode, the prepared film shows a high bioelectrocatalytic activity for glucose oxidation (ca. 15 mA cm?2 at 25 °C) with an extremely high electron‐transfer turnover rate (ca. 650 s?1) comparable to the value for GOD solutions, indicating almost every enzyme molecule entrapped within the ensemble (ca. 3 × 1012 enzymes in a 1 mm × 1 mm film) can work to the fullest extent. This free‐standing, flexible composite film can be used by winding on a needle device; as an example, a self‐powered sugar monitor is demonstrated.  相似文献   

12.
We have previously shown that methionine–heme iron coordination is perturbed in domain-swapped dimeric horse cytochrome c. To gain insight into the effect of methionine dissociation in dimeric cytochrome c, we investigated its interaction with cyanide ion. We found that the Soret and Q bands of oxidized dimeric cytochrome c at 406.5 and 529 nm redshift to 413 and 536 nm, respectively, on addition of 1 mM cyanide ion. The binding constant of dimeric cytochrome c and cyanide ion was obtained as 2.5 × 104 M?1. The Fe–CN and C–N stretching (ν Fe–CN and ν CN) resonance Raman bands of CN?-bound dimeric cytochrome c were observed at 443 and 2,126 cm?1, respectively. The ν Fe–CN frequency of dimeric cytochrome c was relatively low compared with that of other CN?-bound heme proteins, and a relatively strong coupling between the Fe–C–N bending and porphyrin vibrations was observed in the 350–450-cm?1 region. The low ν Fe–CN frequency suggests weaker binding of the cyanide ion to dimeric cytochrome c compared with other heme proteins possessing a distal heme cavity. Although the secondary structure of dimeric cytochrome c did not change on addition of cyanide ion according to circular dichroism measurements, the dimer dissociation rate at 45 °C increased from (8.9 ± 0.7) × 10?6 to (3.8 ± 0.2) × 10?5 s?1, with a decrease of about 2 °C in its dissociation temperature obtained with differential scanning calorimetry. The results show that diatomic ligands may bind to the heme iron of dimeric cytochrome c and affect its stability.  相似文献   

13.
Experiments were designed to test for functional differences which might shed light on the differences in actin-activated ATPase activities recently reported for myosin subfragments-1 bearing different light chains. By using the method of A. G. Weeds and R. S. Taylor (1975, Nature (London)257, 54), two types of subfragment-1 (S-1) from myosin of rabbit fast skeletal muscle were prepared: (S-1)·A1 and (S-1)·A2 bearing, respectively, the alkali-1 and alkali-2 light chains. (In agreement with the findings of these investigators, actin enhanced the ATPase activity of (S-1)·A1 more than that of (S-1)·A2 at lower actin concentrations.) Through use of time-resolved fluorescence depolarization techniques, the affinity constants for the binding of the two types of S-1 to F-actin in the absence of ATP were found to be very similar: 3.4 ± 0.3 × 106m?1 (N = 10) for (S-1)·A1 and 3.9 ± 0.2 × 106m?1 (N = 7) for (S-1)·A2 of one preparation, and 6.4 ± 0.2 × 106m?1 (N = 6) for (S-1)·A1 and 7.7 ± 0.5 × 106m?1 (N = 12) for (S-1)·A2 of another preparation (pH 7.0, 25 °C, 0.28 m KCl, 1.5 mm MgCl2, 0.5 mm ethylene glycol bis (β-aminoethyl ether) N,N′-tetracetic acid, 10 mm imidazole, and 0.1 mmN-tris (hydroxymethyl) methyl-2-aminoethane sulfonate). The affinity constants for the two species of S-1 and actin also have a similar dependence on ionic strength and are not affected by addition of 0.6 mm CaCl2 to the above solution. The CaATPase (or the CaITPase) activities of the two species of S-1 show the same pH dependence.  相似文献   

14.
In order to characterize the vertical variation of abundance and community composition of ammonia-oxidizing archaea (AOA) and ammonia-oxidizing bacteria (AOB) in sediments of a eutrophic lake, Lake Taihu, molecular techniques including real-time PCR, clone library, and sequencing were carried out in this study. Abundances of archaeal amoA gene (ranged from 2.34 × 106 to 4.43 × 107 copies [g dry sediment]?1) were higher than those of bacterial amoA gene (ranged from 5.02 × 104 to 6.91 × 106 copies [g dry sediment]?1) for all samples and both of them exhibited negative correlations with the increased depths. Diversities of archaeal and bacterial amoA gene increased with the elevated depths. There were no significant variations of AOB community structures derived from different sediment depths, whereas obvious differences were observed for the AOA community compositions. The information acquired in this study would be useful to elucidate the roles of AOA and AOB in the nitrogen cycling of freshwater ecosystems.  相似文献   

15.
The uptake of 14C glucose by natural microbial populations has been studied in the Severn Estuary and Bristol Channel, U.K.; the turbidity (suspended solids) in the estuary varied between < 5 mg · 1?1 at the seaward extremity to >800 mg · 1?1 in the estuary proper. The heterotrophic potential, Vm, was found to correlate with turbidity and particulate organic carbon but there was no correlation between microbial biomass, as assessed by plate counts, and turbidity or Vm; measurement of Vm ranged from 0.9 × 10?4 to 288 × 10?4μgC·1?1·h?1 and turnover time from <2 to >100 h. In 17 out of 42 experiments, the uptake of 14C glucose did not conform to Michaelis kinetics and in five of these experiments the data suggested that there may be a threshold of glucose concentration below which there is no uptake.  相似文献   

16.
The electron transfer reactions of horse heart cytochrome c with a series of amino acid-pentacyanoferrate(II) complexes have been studied by the stopped-flow technique, at 25°C, μ = 0.100, pH 7 (phosphate buffer). A second-order behavior was observed in the case of the Fe(CN)5 (histidine)3? complex, with k = 2.8 x 105 M?1 sec?1. For the Fe(CN)5 (alanine)4? and Fe(CN)5(L-glutamate)5? complexes, only a minor deviation of the second-order behavior, close to the experimental error (k = 3.2 × 105 and 1.6 x 105 M?1 sec?1, respectively) was noted at high concentrations of the reactants (e.g., 6 × 10?4 M). The results are in accord with recent work on the Fe(CN)64?/cytochrome c system demonstrating weak association of the reactants. The calculated self-exchange rate constants including electrostatic interactions for the imidazole,L -histidine, 4-aminopyridine, glycinate, β-alaninate, andL-glutamate pentacyanoferrate(II) complexes were 3.3 × 105, 3.3 × 105, 2.8 × 106,4.1 × 102,5.5 × 102, and 6.0 M?1 sec?1, respectively. Marcus theory calculations for the cytochrome c reactions were interpreted in terms of two nonequivalent binding sites for the complexes, with the metalloprotein self-exchange rate constants varying from 104 M?1 sec?1 (histidine, imidazole, and 4-aminopyridine complexes) to 106 M?1 sec ?1 (glycinate, β-alaninate, and L-glutamate complexes).  相似文献   

17.
NADPH-cytochrome c reductase has been isolated from a top-fermenting ale yeast, Saccharomyces cerevisiae (Narragansett strain), after ca. a 240-fold purification over the initial extract of an acetone powder, with a final specific activity (at pH 7.6, 30 °C) of ca. 150 μmol cytochrome c reduced min?1mg?1 protein. The preparation appears to be homogeneous by the criteria of: sedimentation velocity; electrophoresis on cellulose acetate in buffers above neutrality; and by polyacrylamide gel electrophoresis. Although the reductase appeared to partially separate into species “A” and “B” on DEAE-cellulose at pH 8.8, the two species have proven to be indistinguishable electrophoretically (above pH 8) and by sedimentation. By sedimentation equilibrium at 20 °C, a molecular weight of ca. 6.8 (± 0.4) × 104 was obtained with use of a V?20 ° = 0.741 calculated from its amino acid composition. After disruption in 4 m guanidinium chloride- 10 mm dithioerythritol- 1 mm EDTA, pH 6.4 at 20 °C, an M?r of 3.4 (± 0.1) × 104 resulted, which points to a subunit structure of two polypeptide chains per mole. Confirmatory evidence of the two-subunit structure with similar, if not identical, polypeptide chains was obtained by polyacrylamide gel electrophoresis in dodecyl-sulfate, after disruption in 4 m urea and 2% sodium dodecyl sulfate, and yielded a subunit molecular weight of ca. 4 × 104. Sulfhydryl group titration with 4,4′-dithiodipyridine under acidic conditions revealed one sulfhydryl group per monomer, which apparently is necessary for the catalytic reduction of cytochrome c. NADPH, as well as FAD, protects this-SH group from reaction with 5,5′-dithiobis (2-nitrobenzoate). The visible absorption spectrum of the oxidized enzyme (as prepared) has absorption maxima at 383 and 455 nm, typical of a flavoprotein. Flavin analysis (after dissociation by thermal denaturation of the “A” protein) conducted fluorometrically, revealed the presence of 2.0 mol of FAD per 70,000 g, in confirmation of the deduced subunit structure. The identity of the FAD dissociated from either “A” or “B” protein was confirmed by recombination with apo-d-amino acid oxidase and by thin-layer chromatography. A kinetic approach was used to estimate the dissociation constant for either FAD or FMN (which also yields a catalytically active enzyme) to the apoprotein reductase at 30 °C and pH 7.6 (0.05 m phosphate) and yielded values of 4.7 × 10?8m for FAD and 4.4 × 10?8m for FMN.  相似文献   

18.
We estimated the grazing impact of the heterotrophic flagellate Collodictyon triciliatum on the harmful, bloom-forming cyanobacterium Microcystis aeruginosa in an experimental pond during a Microcystis bloom from summer to winter in 2010. For these experiments, we calculated the grazing rates from the digestion rate of C. triciliatum and its food vacuole contents. During the study period, M. aeruginosa exhibited one bloom event with a maximum density of 1.1 × 105 cells ml?1. The cell density of C. triciliatum fluctuated from below the detection limit to 291 cells ml?1. The number of M. aeruginosa cells ingested by C. triciliatum food vacuoles ranged between 0.4 and 10.8 cells flagellate?1, and the digestion rate of C. triciliatum at 25 °C was 0.73 % cell contents min?1. The grazing rate of C. triciliatum on the M. aeruginosa prey was 0.2–6.9 cells flagellate?1 h?1, and its grazing impact was 0.0–25.3 % standing stock day?1. The functional response of C. triciliatum to the M. aeruginosa prey followed the Michaelis–Menten model of significance (r 2 = 0.873, p < 0.001) in our experimental systems, in which the prey concentration varied from 1.0 × 104 to 2.1 × 106 cells ml?1. The maximum grazing rate was 6.2 prey cells grazer?1 h?1, and the half-saturation constant was 1.2 × 105 cells ml?1. We present evidence that C. triciliatum grazing explained the remarkable decrease in M. aeruginosa cell density in the pond. The present study is the first demonstration of the high potential of protistan grazing on M. aeruginosa to reduce cyanobacterial blooms.  相似文献   

19.
The use of microalgae for biofuel production has the potential to reduce fossil fuel consumption. Ideal candidate species of microalgae for bio-oil production need both relatively high growth rates and lipid content. Here, we report on the effects of temperature, nutrients (N, Si), and salinity on growth rates and lipid content of the common freshwater diatom, Fragilaria capucina (Desm), isolated from western Lake Erie. At low NaCl salinity, growth rate increased rapidly from 10 to 20°C, and then further increased slowly from 20 to 30°C, with a maximum specific growth rate of 0.61?day?1. Growth rate declined with increasing salinity (e.g., reduced by ca. 50 and 100% at 137 and 274?mmol?L?1 NaCl, respectively), and increased with increased N and Si concentration until ca. 100?μmol?L?1 for each (with >85% of maximum growth rate at 10?μmol?L?1). Lipid content (% total lipid per dry mass) in nutrient-replete cultures was 14% and (1) increased to >30% at low N and, especially, low Si; (2) was lower at 30°C vs. 20 or 10°C; and (3) decreased with salinity. Thus, F. capucina accumulates lipid to high levels even under N, Si, and temperature levels that permit a high growth rate for this species, and hence, this species is a candidate for use in biofuel production.  相似文献   

20.
The fungicide triarimol was tested for its effect on abscisic acid (ABA) accumulation in growing culturesof Cercospora rosicola. ABA accumulation was reduced by approximately 50% with 10?8 M triarimol. Growth ofC. rosicola, as measured by dry weight accumulation, was inhibited by triarimol concentrations at or greater than 10?7 M. These results are compared with those obtained with clomazone, ancymidol, and paclobutrazol, which inhibit ABA accumulation by 50% at concentrations of 5 × 10?5, 5 × 10?6, and 5 × 10?7 M, respectively. Triarimol, therefore, is among the most potent inhibitors of ABA biosynthesis reported to date. Feeding studies with [14C]mevalonic acid confirmed the inhibition of ABA biosynthesis by 5 × 10?8 M triarimol. These results support previous suggestions that one or more of the steps in the ABA biosynthetic pathway from mevalonic acid is catalyzed by cytochrome P-450. Feeding studies with 1′-deoxy-[2H]-ABA in resuspended cultures ofC. rosicola show that the conversion of this substrate is not inhibited by triarimol.  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