首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The oxidation of proteins and other macromolecules by radical species under conditions of oxidative stress can be modulated by antioxidant compounds. Decreased levels of the antioxidants glutathione and ascorbate have been documented in oxidative stress-related diseases. A radical generated on the surface of a protein can: (1) be immediately and fully repaired by direct reaction with an antioxidant; (2) react with dioxygen to form the corresponding peroxyl radical; or (3) undergo intramolecular long range electron transfer to relocate the free electron to another amino acid residue. In pulse radiolysis studies, in vitro production of the initial radical on a protein is conveniently made at a tryptophan residue, and electron transfer often leads ultimately to residence of the unpaired electron on a tyrosine residue. We review here the kinetics data for reactions of the antioxidants glutathione, selenocysteine, and ascorbate with tryptophanyl and tyrosyl radicals as free amino acids in model compounds and proteins. Glutathione repairs a tryptophanyl radical in lysozyme with a rate constant of (1.05 ± 0.05) × 105 M–1 s–1, while ascorbate repairs tryptophanyl and tyrosyl radicals ca. 3 orders of magnitude faster. The in vitro reaction of glutathione with these radicals is too slow to prevent formation of peroxyl radicals, which become reduced by glutathione to hydroperoxides; the resulting glutathione thiyl radical is capable of further radical generation by hydrogen abstraction. Although physiologically not significant, selenoglutathione reduces tyrosyl radicals as fast as ascorbate. The reaction of protein radicals formed on insulin, β-lactoglobulin, pepsin, chymotrypsin and bovine serum albumin with ascorbate is relatively rapid, competes with the reaction with dioxygen, and the relatively innocuous ascorbyl radical is formed. On the basis of these kinetics data, we suggest that reductive repair of protein radicals may contribute to the well-documented depletion of ascorbate in living organisms subjected to oxidative stress.  相似文献   

2.
The aim of this study was to determine if athletes with a history of hamstring strain injury display lower levels of surface EMG (sEMG) activity and median power frequency in the previously injured hamstring muscle during maximal voluntary contractions. Recreational athletes were recruited, 13 with a history of unilateral hamstring strain injury and 15 without prior injury. All athletes undertook isokinetic dynamometry testing of the knee flexors and sEMG assessment of the biceps femoris long head (BF) and medial hamstrings (MHs) during concentric and eccentric contractions at ±180 and ±60° s?1. The knee flexors on the previously injured limb were weaker at all contraction speeds compared to the uninjured limb (+180° s?1 p = 0.0036; +60° s?1 p = 0.0013; ?60° s?1 p = 0.0007; ?180° s?1 p = 0.0007) whilst sEMG activity was only lower in the BF during eccentric contractions (?60° s?1 p = 0.0025; ?180° s?1 p = 0.0003). There were no between limb differences in MH sEMG activity or median power frequency from either BF or MH in the injured group. The uninjured group showed no between limb differences in any of the tested variables. Secondary analysis comparing the between limb difference in the injured and the uninjured groups, confirmed that previously injured hamstrings were mostly weaker (+180° s?1 p = 0.2208; +60° s?1 p = 0.0379; ?60° ?1 p = 0.0312; ?180° s?1 p = 0.0110) and that deficits in sEMG were confined to the BF during eccentric contractions (?60° s?1 p = 0.0542; ?180° s?1 p = 0.0473). Previously injured hamstrings were weaker and BF sEMG activity was lower than the contralateral uninjured hamstring. This has implications for hamstring strain injury prevention and rehabilitation which should consider altered neural function following hamstring strain injury.  相似文献   

3.
The impact of flow velocity on initial ciliate colonization dynamics on surfaces were studied in the third order Ilm stream (Thuringia, Germany) at a slow flowing site (0.09 m s?1) and two faster flowing sites (0.31 m s?1) and in flow channels at 0.05, 0.4, and 0.8 m s?1. At the slow flowing stream site, surfaces were rapidly colonized by ciliates with up to 60 cells cm?2 after 24 h. In flow channels, the majority of suspended ciliates and inorganic matter accumulated at the surface within 4.5 h at 0.05 m s?1. At 0.4 m s?1 the increase in ciliate abundance in the biofilm was highest between 72 and 168 h at about 3 cells cm?2 h?1. Faster flow velocities were tolerated by vagile flattened ciliates that live in close contact to the surface. Vagile flattened and round filter feeders preferred biofilms at slow flow velocities. Addition of inorganic particles (0, 0.6, and 7.3 mg cm?2) did not affect ciliate abundance in flow channel biofilms, but small ciliate species dominated and number of species was lowest (16 species cm?2) in biofilms at high sediment content. Although different morphotypes dominated the communities at contrasting flow velocities, all functional groups contributed to initial biofilm communities implementing all trophic links within the microbial loop.  相似文献   

4.
The action of C60 fullerene and its derivatives as a radical-scavenging antioxidant has received much attention, but their reactivity toward free radicals and antioxidant capacity have not been well elucidated yet. In the present study, the reactivity of the two types of water-soluble, sugar-pendant C60 fullerenes, C60-1S and C60-2S, toward peroxyl radical and their effect against human plasma lipid peroxidation were measured. The rate constants for the reaction of C60-1S and C60-2S with peroxyl radicals were obtained from their effect on the bleaching of β-carotene in lipid-SDS micelle system as 4.6 × 103 and 8.0 × 103 M?1 s?1 at 37 °C, respectively. They inhibited the free radical-induced lipid peroxidation in human plasma in a concentration-dependent manner. These results suggest that the sugar-pendant fullerenes C60-1S and C60-2S act as a radical-scavenging antioxidant with the activity similar to the phenolic antioxidants.  相似文献   

5.
Phosphopantothenoylcysteine synthetase (PPCS) catalyzes the formation of phosphopantothenoylcysteine from (R)-phosphopantothenate and l-cysteine with the concomitant consumption of a nucleotide triphosphate. Herein, the human coaB gene encoding PPCS is cloned into pET23a and overexpressed in E. coli BL21(DE3), to yield 10 mg of purified enzyme per liter of culture. Detailed kinetic studies found that this PPCS follows a similar Bi Uni Uni Bi Ping Pong mechanism as previously described for the E. faecalis PPCS, except that the human enzyme can use both ATP and CTP with similar affinity. One significant difference for human PPCS catalysis with respect to ATP and CTP is that the enzyme shows cooperative binding of ATP, measured as a Hill constant of 1.7. PPCS catalysis under CTP conditions displayed Michaelis constants of 265 μM, 57 μM, and 16 μM for CTP, PPA, and cysteine, respectively, with a kcat of 0.53 ± 0.01 s? 1 for the reaction. Taking into account the cooperativity under ATP condition, PPCS exhibited Michaelis constants of 269 μM, 13 μM, and 14 μM for ATP, PPA, and cysteine, respectively, with a kcat of 0.56 s? 1 for the reaction. Oxygen transfer studies found that 18O from [carboxyl-18O] phosphopantothenate is incorporated into the AMP or CMP produced during PPCS catalysis, consistent with the formation of a phosphopantothenoyl cytidylate or phosphopantothenoyl adenylate intermediate, supporting similar catalytic mechanisms under both CTP and ATP conditions. Inhibition studies with GTP and UTP as well as product inhibition studies with CMP and AMP suggest that human PPCS lacks strong nucleotide selectivity.  相似文献   

6.
The two components (BinA and BinB) of Lysinibacillus sphaericus binary toxin together are highly toxic to Culex and Anopheles mosquito larvae, and have been employed world-wide to control mosquito borne diseases. Upon binding to the membrane receptor an oligomeric form (BinA2.BinB2) of the binary toxin is expected to play role in pore formation. It is not clear if these two proteins interact in solution as well, in the absence of receptor. The interactions between active forms of BinA and BinB polypeptides were probed in solution using size-exclusion chromatography, pull-down assay, surface plasmon resonance, circular dichroism, and by chemically crosslinking BinA and BinB components. We demonstrate that the two proteins interact weakly with first association and dissociation rate constants of 4.5 × 103 M?1 s?1 and 0.8 s?1, resulting in conformational change, most likely, in toxic BinA protein that could kinetically favor membrane translocation of the active oligomer. The weak interactions between the two toxin components could be stabilized by glutaraldehyde crosslinking. The cross-linked complex, interestingly, showed maximal Culex larvicidal activity (LC50 value of 1.59 ng mL?1) reported so far for combination of BinA/BinB components, and thus is an attractive option for development of new bio-pesticides for control of mosquito borne vector diseases.  相似文献   

7.
This study is aimed at identifying the proteins that are up-regulated during astaxanthin accumulation in Haematococcus lacustris. For this H. lacustris cells were cultivated in photobioreactors under normal light irradiance of 40 μE m?2 s?1 for 6 days and then induced to accumulate astaxanthin for 3 days further by exposure to continuous high irradiance of 200 μE m?2 s?1 with fluorescent lamps as light source after the cells reached the stationary phase in a nitrogen-depleted condition. Under this condition, the average astaxanthin content per cell increased from 91 mg/l up to 406 mg/l after 3 days of induction. The proteomics data from a two-dimensional electrophoretic comparison demonstrated that a combination of nitrogen source depletion and 1 h high light have significantly changed the pattern of protein expression in H. lacustris. A total of 49 protein spots were picked after 1 h of stress induction. They consisted of 13 down-regulated proteins and 36 up-regulated proteins. Fifteen proteins which had highly up-regulated expression were further analyzed by matrix assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF-MS). The results will point toward interesting proteins that can be pursued for further analysis of astaxanthin biosynthesis pathway.  相似文献   

8.
A β-carbonic anhydrase (CA, EC 4.2.1.1), the protein encoded by the NCE103 gene of Candida glabrata which also present in Candida albicans and Saccharomyces cerevisiae, was cloned, purified, characterized kinetically and investigated for its inhibition by a series simple, inorganic anions such as halogenides, pseudohalogenides, bicarbonate, carbonate, nitrate, nitrite, hydrogen sulfide, bisulfite, perchlorate, sulfate and some isosteric species. The enzyme showed significant CO2 hydrase activity, with a kcat of 3.8 × 105 s?1 and kcat/KM of 4.8 × 107 M?1 s?1. The Cà glabrata CA (CgCA) was moderately inhibited by metal poisons (cyanide, azide, cyanate, thiocyanate, KIs of 0.60–1.12 mM) but strongly inhibited by bicarbonate, nitrate, nitrite and phenylarsonic acid (KIs of 86–98 μM). The other anions investigated showed inhibition constants in the low millimolar range, with the exception of bromide and iodide (KIs of 27–42 mM).  相似文献   

9.
The phytase of Sporotrichum thermophile was purified to homogeneity using acetone precipitation followed by ion-exchange and gel-filtration column chromatography. The purified phytase is a homopentamer with a molecular mass of ~456 kDa and pI of 4.9. It is a glycoprotein with about 14% carbohydrate, and optimally active at pH 5.0 and 60 °C with a T1/2 of 16 h at 60 °C and 1.5 h at 80 °C. The activation energy of the enzyme reaction is 48.6 KJ mol?1 with a temperature quotient of 1.66, and it displayed broad substrate specificity. Mg2+ exhibited a slight stimulatory effect on the enzyme activity, while it was markedly inhibited by 2,3-butanedione suggesting a possible role of arginine in its catalysis. The chaotropic agents such as guanidinium hydrochloride, urea and potassium iodide strongly inhibited phytase activity. Inorganic phosphate inhibited enzyme activity beyond 3 mM. The maximum hydrolysis rate (Vmax) and apparent Michaelis–Menten constant (Km) for sodium phytate were 83 nmol mg?1 s?1 and 0.156 mM, respectively. The catalytic turnover number (Kcat) and catalytic efficiency (Kcat/Km) of phytase were 37.8 s?1 and 2.4 × 105 M?1 s?1, respectively. Based on the N-terminal and MALDI–LC–MS/MS identified amino acid sequences of the peptides, the enzyme did not show a significant homology with the known phytases.  相似文献   

10.
Light is one of the most important environmental signals regulating physiological processes of many microorganisms. However, very few studies have been reported on the qualitative or quantitative effects of light on control of postharvest spoilage using antagonistic bacteria. In this study, we investigated the effects of white, red, green, and blue light at photon flux densities of 40, 240, and 360 μmol m?2 s?1 on Bacillus amyloliquefaciens JBC36 (JBC36), which has been reported as a promising candidate for biocontrol of green and blue mold on mandarin fruit. With the exception of blue light at 240 and 360 μmol m?2 s?1, light generally stimulated growth of JBC36 compared to the controls grown in the dark. Red light increased swarming motility irrespective of intensity and significantly enhanced biofilm formation at 240 μmol m?2 s?1. Production of antifungal metabolites and antifungal activity on Penicillium digitatum was also affected by light quality. Interestingly, antifungal activity was significantly increased when JBC36 and P. digitatum was co-incubated under red and green light at an intensity of 240 μmol m?2 s?1. We also demonstrated that the quality of light resulted in changes in colonization of JBC36 on mandarin fruit and control of green mold. In particular, red light increased the population level on mandarin fruit and biocontrol efficacy against green mold. These results represent the first report on the effect of light quality on an antagonistic bacterium for the control of postharvest spoilage. We believe that an improved understanding of the JBC36 response to light quality may help in the development of strategies to increase biocontrol efficacy of postharvest spoilage.  相似文献   

11.
Mandelate racemase (MR, E.C. 5.1.2.2) from Pseudomonas putida catalyzes the Mg2+-dependent 1,1-proton transfer that interconverts the enantiomers of mandelate and has been studied extensively as a model for understanding how enzymes catalyze the deprotonation of carbon acid substrates with relatively high pKa values. Purification of recombinant MR as a fusion protein with an N-terminal hexahistidine tag using immobilized-nickel ion affinity chromatography and elution with a linear gradient of EDTA revealed three enzyme species (mrI, mrII, and mrIII). While mrIII was catalytically inactive, both mrI and mrII catalyzed the racemization of (S)-mandelate with turnover numbers (kcat) of 190 ± 22 and 940 ± 24 s?1, respectively. Circular dichroism analysis suggested that mrIII was a partially unfolded or misfolded form of the enzyme. Replacement of the N-terminal hexahistidine tag by a StrepII-tag appeared to ameliorate the folding problem yielding a single enzyme species with a turnover number of 1124 ± 43 s?1. The MR fusion protein bearing an N-terminal StrepII-tag and a C-terminal decahistidine tag also exhibited reduced turnover (kcat = 472 ± 37 s?1). These results highlight a potential problem that may be encountered when producing fusion enzymes bearing a polyhistidine tag: soluble, active enzyme may be obtained but care must be taken to ensure that it is free of minor misfolded forms that can alter the apparent activity of the enzyme.  相似文献   

12.
The antioxidant activities of isoorientin-6″-O-glucoside were studied using various models. Isoorientin-6″-O-glucoside was more potent than Trolox, probucol and butylated hydroxytoluene (BHT) in reducing the stable free radical 1,1-diphenyl-2-picrylhydrazyl (DPPH). It also scavenged superoxide anion, peroxyl and hydroxyl radicals that were generated by xanthine/xanthine oxidase, 2,2′-azobis(2-amidinopropane) dihydrochloride (AAPH) and Fe3+–ascorbate–EDTA–H2O2 system, respectively. The IC50 value, stoichiometry factor and second-order rate constant were 9.0 ± 0.8 μM, 1.8 ± 0.1 and 2.6 × 1010 M−1 s−1 for superoxide generation, peroxyl and hydroxyl radicals. However, isoorientin-6″-O-glucoside did not inhibit xanthine oxidase activity or scavenge hydrogen peroxide (H2O2), carbon radical or 2,2′-azobis(2,4-dimethylvaleronitrile) (AMVN)-derived peroxyl radical in hexane. Isoorientin-6″-O-glucoside inhibited Cu2+-induced oxidation of human low-density lipoprotein (LDL) as measured by fluorescence intensity, thiobarbituric acid-reactive substance formation and electrophoretic mobility. Since isoorientin-6″-O-glucoside did not possess pro-oxidant activity, it may be an effective water-soluble antioxidant that can prevent LDL against oxidation.  相似文献   

13.
Hypochlorous acid and its acid–base counterpart, hypochlorite ions, produced under inflammatory conditions, may produce chloramides of glycosaminoglycans, these being significant components of the extracellular matrix (ECM). This may occur through the binding of myeloperoxidase directly to the glycosaminoglycans. The N–Cl group in the chloramides is a potential selective target for both reducing and oxidizing radicals, leading possibly to more efficient and damaging fragmentation of these biopolymers relative to the parent glycosaminoglycans. In this study, the fast reaction techniques of pulse radiolysis and nanosecond laser flash photolysis have been used to generate both oxidizing and reducing radicals to react with the chloramides of hyaluronan (HACl) and heparin (HepCl). The strong reducing formate radicals and hydrated electrons were found to react rapidly with both HACl and HepCl with rate constants of 1–1.7×108 and 0.7–1.2×108 M−1 s−1 for formate radicals and 2.2×109 and 7.2×108 M−1 s−1 for hydrated electrons, respectively. The spectral characteristics of the products of these reactions were identical and were consistent with initial attack at the N–Cl groups, followed by elimination of chloride ions to produce nitrogen-centered radicals, which rearrange subsequently and rapidly to produce C-2 radicals on the glucosamine moiety, supporting an earlier EPR study by M.D. Rees et al. (J. Am. Chem. Soc. 125: 13719–13733; 2003). The oxidizing hydroxyl radicals also reacted rapidly with HACl and HepCl with rate constants of 2.2×108 and 1.6×108 M−1 s−1, with no evidence from these data for any degree of selective attack on the N–Cl group relative to the N–H groups and other sites of attack. The carbonate anion radicals were much slower with HACl and HepCl than hydroxyl radicals (1.0×105 and 8.0×104 M−1 s−1, respectively) but significantly faster than with the parent molecules (3.5×104 and 5.0×104 M−1 s−1, respectively). These findings suggest that these potential in vivo radicals may react in a site-specific manner with the N–Cl group in the glycosaminoglycan chloramides of the ECM, possibly to produce more efficient fragmentation. This is the first study therefore to conclusively demonstrate that reducing radicals react rapidly with glycosaminoglycan chloramides in a site-specific attack at the N–Cl group, probably to produce a 100% efficient biopolymer fragmentation process. Although less reactive, carbonate radicals, which may be produced in vivo via reactions of peroxynitrite with serum levels of carbon dioxide, also appear to react in a highly site-specific manner at the N–Cl group. It is not yet known if such site-specific attacks by this important in vivo species lead to a more efficient fragmentation of the biopolymers than would be expected for attack by the stronger oxidizing species, the hydroxyl radical. It is clear, however, that the N–Cl group formed under inflammatory conditions in the extracellular matrix does present a more likely target for both reactive oxygen species and reducing species than the N–H groups in the parent glycosaminoglycans.  相似文献   

14.
The objective of this study was to investigate nitrification rates in algal–bacterial biofilms of waste stabilization ponds (WSP) under different conditions of light, oxygen and pH. Biofilms were grown on wooden plates of 6.0 cm by 8.0 cm by 0.4 cm in a PVC tray continuously fed with synthetic wastewater with initial NH4-N and Chemical Oxygen Demand (COD) concentrations of 40 mg l?1 and 100 mg l?1, respectively, under light intensity of 85–95 μE m?2 s?1. Batch activity tests were carried out by exposure of the plates to light conditions as above (to simulate day time), dim light of 1.8–2.2 μE m?2 s?1 (to simulate reduced light as in deeper locations in WSP) and dark conditions (to simulate night time). Dissolved oxygen (DO) concentration and pH were controlled. At some experiments, both parameters were kept constant, and at others they were left to vary as in WSP. Results show biofilm nitrification rates of 945–1817 mg-N m?2 d?1 and 1124–1615 mg-N m?2 d?1 for light and dark experiments. When the minimum DO was 4.1 mg l?1, the biofilm nitrification rates under light and dark conditions did not differ significantly at 95% confidence. When the minimum DO in the dim light experiment was 3.2 mg l?1, the nitrification rates under light and dim light conditions were 945 mg-N m?2 d?1 and 563 mg-N m?2 d?1 and these significantly differed. Further decrease of DO to 1.1 mg l?1 under dark conditions resulted in more decrease of the nitrification rates to 156 mg-N m?2 d?1. It therefore seems that under these experimental conditions, biofilm nitrification rates are significantly reduced at a certain point when bulk water DO is between 3.2 mg l?1 and 4.1 mg l?1. As long as bulk water DO under dark is high, light is not important in influencing the process of nitrification.  相似文献   

15.
Meriem Alami  Dusan Lazar  Beverley R. Green 《BBA》2012,1817(9):1557-1564
Aureococcus anophagefferens is a picoplanktonic microalga that is very well adapted to growth at low nutrient and low light levels, causing devastating blooms (“brown tides”) in estuarine waters. To study the factors involved in long-term acclimation to different light intensities, cells were acclimated for a number of generations to growth under low light (20 μmol photons m? 2 s? 1), medium light (60 or 90 μmol photons m? 2 s? 1) and high light (200 μmol photons m? 2 s? 1), and were analyzed for their contents of xanthophyll cycle carotenoids (the D pool), fucoxanthin and its derivatives (the F pool), Chls c2 and c3, and fucoxanthin Chl a/c polypeptides (FCPs). Higher growth light intensities resulted in increased steady state levels of both diadinoxanthin and diatoxanthin. However, it also resulted in the conversion of a significant fraction of fucoxanthin to 19′-butanoyloxyfucoxanthin without a change in the total F pool. The increase in 19′-butanoyloxyfucoxanthin was paralleled by a decrease in the effective antenna size, determined from the slope of the change in F0 as a function of increasing light intensity. Transfer of acclimated cultures to a higher light intensity showed that the conversion of fucoxanthin to its derivative was a relatively slow process (time-frame of hours). We suggest the replacement of fucoxanthin with the bulkier 19′-butanoyloxyfucoxanthin results in a decrease in the light-harvesting efficiency of the FCP antenna and is part of the long-term acclimative response to growth at higher light intensities.  相似文献   

16.
A thermostable adenylate kinase (tAK) has been used as model protein contaminant on surfaces, so used because residual protein after high temperature wash steps can be detected at extremely low concentrations. This gives the potential for accurate, quantitative measurement of the effectiveness of different wash processes in removing protein contamination. Current methods utilise non-covalent (physisorbtion) of tAK to surfaces, but this can be relatively easily removed. In this study, the covalent binding of tAK to surfaces was studied to provide an alternative model for surface contamination. Kinetic analysis showed that the efficiency of the enzyme expressed as the catalytic rate over the Michaelis constant (kcat/KM) increased from 8.45±3.04 mM?1 s?1 in solution to 32.23±3.20 or 24.46±4.41 mM?1 s?1 when the enzyme was immobilised onto polypropylene or plasma activated polypropylene respectively. Maleic anhydride plasma activated polypropylene showed potential to provide a more robust challenge for washing processes as it retained significantly higher amounts of tAK enzyme than polypropylene in simple washing experiments. Inhibition of the coupled enzyme (luciferase/luciferin) system used for the detection of adenylate kinase activity, was observed for a secondary product of the reaction. This needs to be taken into consideration when using the assay to estimate cleaning efficacy.  相似文献   

17.
Thermal stability of starch degrading enzymes varies from one source to another. This research was aimed to study thermal stability of starch degrading enzymes of teff malt. Isothermal mashing at temperatures ranging between 40 and 75 °C with sampling in 15 min interval for a total of 90 min was conducted. The study showed that deactivation rate constants of alpha- and beta-amylases ranged from 0.0003 to 0.0409 min?1, and 0.002 to 0.032 min?1, respectively. Rate of deactivation of limit dextrinase was not significant at temperatures lower than 60 °C but showed high deactivation at higher temperatures with rate constants ranging from 0.02 to 0.1 min?1. The thermal deactivation energies of alpha-amylase, beta-amylase, and limit dextrinase were found to be 148, 82, and 144 kJ/mol, respectively. The present findings have significant applications in commercial processes where determination of the upper temperature limits for these enzymes is required.  相似文献   

18.
The study had three purposes: to verify a hypothesized speed decrease during the 15 km cross-country sit skiing (CCSS) race; documenting this possible fatigue effect (speed decrease), to evaluate changes among the four laps in kinematics parameters (cycle speed, cycle duration, cycle length, duty cycle (percentage ratio between pushing and total cycle duration), pole inclination, trunk inclination and shoulder–hand distance); to compare the kinematics parameters in cross-country sit skiers of different level. Video recordings were carried out during the 2006 Turin Winter Paralympic Games with two conventional digital video-cameras positioned on a flat and an uphill (8.3°) track, respectively. Better performing skiers (G1) had significantly higher speeds than worse performers (G2) both in the flat (6.54 ± 0.13 vs. 5.89 ± 0.50 m s?1 and 5.55 ± 0.14 vs. 4.62 ± 0.22 m s?1 in the first and last lap, respectively) and in the uphill track (3.67 ± 0.45 vs. 3.05 ± 0.59 m s?1 and 3.20 ± 0.36 vs. 2.26 ± 0.36 m s?1 in the first and last lap, respectively). The G1 athletes were able to maintain the high-speed better than the G2 over the entire race. Significant differences in cycle length and duty cycle between groups would be justified by the higher physical fitness of G1 skiers.  相似文献   

19.
Nitrate removal rates in woodchip media of varying age   总被引:1,自引:0,他引:1  
A variety of low-cost carbonaceous solids have been successfully tested in bioreactors designed for nitrate treatment. In many agricultural and wastewater settings, however, such reactors may be practical only if they are maintenance free for a number of years after installation. Although field installations have demonstrated consistent treatment over multi-year timeframes, the ability to accurately quantify slowly declining reaction rates in field settings is problematic because of variations in reactor flow rates, ambient temperatures and influent chemistry. In this study, laboratory column tests were undertaken on four samples of coarse wood particle media (woodchips), two that were fresh and two that had been in continuous operation in subsurface denitrifying bioreactors for periods of 2 and 7 years respectively. Four experimental runs were undertaken at increasing influent NO3-N concentrations of from 3.1 to 48.8 mg N L?1. Nitrate mass removal rates remained relatively constant and did not systematically increase in successive runs at higher NO3 concentrations indicating that NO3 was not the rate-limiting substrate at these concentrations. Thus, zero-order reaction kinetics were used to model the attenuation reaction (presumably denitrification). The 7-year-old media had a mean NO3-N removal rate of 9.1 mg N L?1 d?1 (6.4 g N m?3 media d?1), which remained within 75% of the rate for the 2-year-old media (12.1 mg N L?1 d?1 or 8.5 g N m?3 media d?11) and within 40–59% of the rate for the fresh chips (15.4–23.0 mg N L?1 d?1 or 10.8–16.1 g N m?3 media d?1). Results support field experience indicating that woodchips loose about 50% of their reactivity during their first year of operation as soluble organic compounds are leached out, but then relatively stable rates persist for a considerable number of years thereafter.  相似文献   

20.
A novel endo-β-1,4-glucanase (EG)-producing strain was isolated and identified as Penicillium purpurogenum KJS506 based on its morphology and internal transcribed spacer (ITS) rDNA gene sequence. P. purpurogenum produced one of the highest levels of EG (5.6 U mg-protein?1) with rice straw and corn steep powder as carbon and nitrogen sources, respectively. The extracellular EG was purified to homogeneity by sequential chromatography of P. purpurogenum culture supernatants on a DEAE sepharose column, a gel filtration column, and then on a Mono Q column with fast protein liquid chromatography. The purified EG was a monomeric protein with a molecular weight of 37 kDa and showed broad substrate specificity with maximum activity towards lichenan. P. purpurogenum EG showed t1/2 value of 2 h at 70 °C and catalytic efficiency of 118 ml mg?1 s?1, one of the highest levels seen for EG-producing microorganisms. Although EGs have been reported elsewhere, the high catalytic efficiency and thermostability distinguish P. purpurogenum EG.  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