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.
Gas exchange characteristics of three major Louisiana Mississippi River deltaic plain marsh species, Spartina patens (Ait.) Muhl., Spartina altemiflora Lois., and Panicum hemitomon Shult., was studied under controlled environment conditions. The optimum temperature for maximum photosynthesis was ≈ 36 °C for S. patens, 27 °C for S. alterniflora, and 28 °C for rP. hemitomon. Net photosynthesis rates at optimum temperature averaged 20.1 μmol · mt-2 · st-1 in S. patens, 22.8 μmol · m−2 · s−1 in S. alterniflora, and 11.4 μmol · m−2 · s−1 in P. hemitomon. Photosynthetic light saturation occurred ≈720, 530, and 750 μmol · m−2 · s−1 in S. patens, S. alterniflora, and P. hemitomon, respectively. Only S. patens had a midday depression of stomatal conductance, but net photosynthesis was not reduced by the depression. Maximum stomatal conductances were 285 mmol · m−2 · s−1 in S. patens, 238 mmol · m−2 · s−1 in S. alterniflora, and 335 mmol · m−2 · s−1 in P. hemitomon. In contract, net photosynthesis values were lower in P. hemitomon compared with the Spartina species, indicating a greater degree of water use efficiency of photosynthesis for both Spartina species.  相似文献   

2.
Characteristics of photosynthesis and respiration of bladelets were compared between Ecklonia cava Kjellman sporophytes growing in a warmer temperate locality (Tei, Kochi Pref., southern Japan) and in a cooler temperate locality (Nabeta, Shizuoka Pref., central Japan). Photosynthesis and respiration were measured with a differential gas-volumeter (Productmeter). In photosynthesis-light curves at 20°C, the rate of net photosynthesis was almost the same at light intensities lower than 25 μmol m−2 s−1 and the light-saturation occurred at 200–400 μmol m−2s−1 in plants of both localities. The light-saturated net photosynthetic rates were higher in winter and spring than in summer and autumn in both plants. The optimum temperature for net photosynthesis at 400 μmol m−2s−1 was 27°C throughout the year in the Tei plant and 25–27°C in the Nabeta plant. The decrease of net photosynthetic rates in the supraoptimal temperature range up to 29°C was sharper in winter and spring than in summer and autumn in both plants, being smaller in the Tei plant than in the Nabeta plant in all seasons. The dark respiration rate always increased with water temperature rise in both plants. No clear differences were found in the dark respiration rate between Tei and Nabeta plants except that when measured against dry weight, the Tei plant showed a slightly lower rate as compared with the Nabeta plant.  相似文献   

3.
Infrared laser traps (optical tweezers) were used to study laser-induced organelle movements in the marine alga Pyrocystis noctiluca (Dinophyta). These cells are highly suitable for optical micromanipulation due to their large size and extensive vacuole. Experiments were done with plastids held by optical tweezers and moved from the nuclear area into the vacuole. The subsequent retraction movement was analysed for speed. The displaced organelles remained connected to their original position by a thin cytoplasmic strand, often less than 1 μm in diameter. When the organelles were released they rapidly returned at an initial rate of 81.7 ± 7.8 μm . s?1 (overall displacement 50 μm, measured distance 20 μm, 25 °C ± 1 °C, number of cells 22), slowing down with progressive retraction of the connecting strand. The return movement was reduced to 4.2 ± 0.2 μ .s?1 (n = 10) when the organelles were displaced and held for 1 min. Displacement to a longer distance increased the rate of return movement. A change from a high to a low environmental temperature significantly reduced movement from 94.5 ± 9.0 . s?1 (30 °C ± 1 °C, n = 22) to 34.5 ± 2.7 μm .s?1 (5°C ± 1 °C, n = 22). Nocodazole and N-ethylmaleimide (NEM), inhibitors of microtubules and acto-myosin, respectively, did not affect the retraction of the connecting strand, but at high concentrations of NEM it became increasingly difficult to move organelles away from the nuclear area. We suggest that the return movement of organelles within laser-induced artificial strands mainly depends on the viscoelastic properties of the tonoplast. The quantification of these properties by optical tweezers allows determination of reactions of plant cells to temperature changes.  相似文献   

4.
《Bioorganic chemistry》1987,15(2):100-108
Nonenzymatic rates of hydrolytic deamination of adenosine and cytidine by acids and bases analogous to side chains of naturally occurring amino acids are compared with the rates of uncatalyzed deamination in water and with the rates of the hydroxide- and hydrogen ion-catalyzed reactions. For adenosine, hydroxide ion is an effective catalyst, with a second-order rate constant of 7.5 × 10−6 m−1 s−1 at 85°C and an energy of activation of 19.9 kcal/mol. Acid-catalyzed deamination of adenine proceeds with a second-order rate constant of 1.5 × 10−6 m−1 s−1 at 85°C. At concentrations of 1 m and at pH values corresponding to their respective pKa values, dimethylamine, acetate, selenide, imidazole, phosphate, and zinc(II) do not enhance the rate of deamination of adenosine beyond that observed in water, and 2-mercaptoethanol produces only a modest rate enhancement. The uncatalyzed rate of adenosine deamination in water is 8.6 × 10−9 s−1 at 85°C: extrapolation to 37°C and comparison with kcat for rat hepatoma adenosine deaminase yield a rate enhancement by the enzyme of approximately 2 × 1012-fold. 1,6-Dimethyladenosine, the conjugate acid of which has a pKa value much higher than that of adenosine, is not readily deaminated, suggesting that the uncatalyzed deamination of adenosine does not proceed by hydroxide ion attack on the rare protonated form of adenosine, but rather by attack on the neutral species. Deamination of cytidine is catalyzed most effectively by hydroxide ion, with a second-order rate constant of 4.5 × 10−4 m−1 s−1 at 85°C and an energy of activation of 28.5 kcal/mol. The uncatalyzed rate of deamination of cytidine in water, which also exhibits an energy of activation of 28.5 kcal/mol, is 8.8 × 10−8 s−1 at 85°C. Comparison of the rate extrapolated to 25°C with kcat for bacterial cytidine deaminase gives a rate enhancement for the enzyme of 4 × 1011-fold. The C-5 proton of the pyrimidine ring of cytidine does not exchange with solvent during alkaline hydrolysis, suggesting that deamination under these conditions does not involve prior addition of water across the 5,6 double bond.  相似文献   

5.
Rates of net photosynthesis and respiration were determined for Pithophora oedogonia (Mont.) Wittr. acclimatized to 56 combinations of light (7–1200 μE m?2 s?1) and temperature (5–35°C). Conditions for maximum net photosynthesis were estimated to be 26°C and 970 μE m?2 s?1. The rate of net photosyntheses varied considerably with temperature, with the maximum measured value (9.67 mg O2 h?1 g dry wt.?1) occurring at 25°C. Respiration rate increased with temperature and the light received just prior to measurement. The maximum respiration rate (7.05 mg O2 g?1 h?1) occurred at 30°C and 1200 μE m?2 s?1. Exposure of Pithophora to light levels of 600 or 1200 μE m?2 s?1 prior to determination of the respiration rate resulted in significantly elevated levels of oxygen consumption at temperatures ≥ 15°C. The relationship between light, temperature and photosynthesis and respiration were summarized as three-dimensional response surfaces.  相似文献   

6.
We determined the maximum sustained swimming speed (Ucrit), and resting and maximum ventilation rates of the Antarctic fish Pagothenia borchgrevinki at five temperatures between −1°C and 8°C. We also determined resting metabolic rate (VO2) at −1°C, 2°C, and 4°C. Ucrit of P. borchgrevinki was highest at −1°C (2.7±0.1 BL s−1) and rapidly decreased with temperature, representing a thermal performance breadth of only 5°C. This narrow thermal performance supports our prediction that specialisation to the subzero Antarctic marine environment is associated with a physiological trade-off in performance at high temperatures. Resting oxygen consumption and ventilation rate increased by more than 200% across the temperature range, which most likely contribute to the decrease in aerobic swimming capabilities at higher temperatures.  相似文献   

7.
The physical properties of intact type IV collagen from the mouse EHS sarcoma were studied in acid solution using laser light scattering and viscometry. The experimentally observed values of molecular weight, translational diffusion coefficient, particle scattering factor at 175.5° and a wavelength of 633 nm and intrinsic viscosity at 22°C were 532000, 0.66 × 10−7cm2s−1, 0.492 and 74.7 ml/g respectively. Plots of Kc/R0 versus collagen concentration were linear with a slope of approximately 0, indicating that under the conditions studied, type IV collagen molecules do not form supra-molecular aggregates. Experimentally determined translational diffusion coefficients closely approximated the calculated value for a rod-like molecule 424 nm long and 1.5 nm in diameter. Based on this observation, it is concluded that the type IV collagen molecule translates like a bent rigid rod similar to the interstitial collagens. However, the low intrinsic viscosity and larger value of the particle scattering factor for type IV collagen molecules in comparison with the interstitial collagens indicate that type IV collagen is considerably more flexible. Physical measurements on molecules in solution are consistent with a model of the type IV molecule containing numerous flexible bends with bend angles less than 125°. It is concluded that the type IV collagen molecule behaves like a worm-like rod in solution.  相似文献   

8.
Frequent off-flavor events caused by geosmin and 2-methylisoborneol (MIB) have attracted research on the main producers, cyanobacteria. This study evaluated the effects of light and temperature on the odor production of MIB-producing Pseudanabaena sp. Lauterborn and geosmin-producing Anabaena ucrainica (Schhorb.) Watanabe. The maximum MIB production and lowest growth rate (indicated by the chlorophyll a (Chl a)) were observed at 35 °C compared with that at 10 °C and 25 °C. Cultures grown under a light intensity of 60 μmol photons m−2 s−1 demonstrated the highest MIB production and minimum growth rate, whereas the minimum MIB production and maximum growth rate were obtained under 10 μmol photons m−2 s−1. Similar patterns were observed for geosmin production. A. ucrainica had the highest geosmin production and lowest Chl a concentration under 10 °C and 60 μmol photons m−2 s−1. Moreover, greater proportions of geosmin and MIB were released into extracellular under growth-inhibiting temperatures and light intensities. An inverse correlation between odor production and the cell growth rate was suggested, and this relationship may reflect the competition for substrates of odor and Chl a synthesis. Thus, the accumulation of geosmin and MIB was probably the result of decreased cellular metabolic activity in cyanobacteria.  相似文献   

9.
In this study, the plant node was dried in an oven (40, 50 and 60 °C), shade and temperature-controlled microwave (40, 50 and 60 °C) methods. Statistically (p<0.05), the values closest to the color values of fresh grass were determined in an oven at 40 °C drying temperature. Effective diffusion values varied between 8.85×10−8–5.65×10−6 m2 s−1. While the activation energy was 61.28 kJ mol−1 in the oven, it was calculated as 85.24 kJ mol−1 in the temperature-controlled microwave. Drying data was best estimated in the Midilli-Küçük (R2 0.9998) model oven at 50 °C. The highest SMER value was calculated as 0.0098 kg kWh−1 in the temperature-controlled microwave drying method. The lowest SEC value in the temperature-controlled microwave was determined as 24.03 kWh kg−1. It was determined that enthalpy values varied between −2484.66/−2623.38 kJ mol−1, entropy values between −162.04/−122.65 J mol−1 and Gibbs free energy values between 453335.22–362581.40 kJ mol−1. Drying rate values were calculated in the range of 0.0127–0.9820 g moisture g dry matter−1 in the temperature-controlled microwave, 0.0003–0.0762 g dry matter−1 in the oven, and 0.001–0.0058 g moisture g dry moisture matter−1 in the shade. Phenolic content 6957.79 μg GAE g−1 fw - 48322.27 μg GAE g−1 dw, flavonoid content 3806.67 mg KE L−1 fw - 22200.00 mg KE L−1 dw and antioxidant capacity 43.35 μmol TE g−1 fw - 323.47 μmol TE g−1 dw. The highest chlorophyll values were obtained from samples dried in an oven at 40 °C. According to the findings, it is recommended to dry the knotweed (Polygonum cognatum Meissn.) plant in a temperature-controlled microwave oven at low temperatures. In this study, in terms of drying kinetics and energy parameters, a temperature-controlled microwave dryer of 60 °C is recommended, while in terms of quality characteristics, oven 40 °C and shade methods are recommended.  相似文献   

10.
The typical musty/earthy odor-causing compound, 2-methylisoborneol (MIB), is usually associated with the occurrence and proliferation of benthic/subsurface-living cyanobacteria in source water. Control of MIB-producing cyanobacteria in source water may greatly reduce the processing burden for drinking water treatment plants. We explored the mechanism and feasibility of restricting the growth of subsurface-living Planktothrix sp. by reducing underwater light availability. The effects of light intensity (5, 17, 36, 85, and 250 μmol photons m−2 s−1) on the growth and MIB production of Planktothrix sp. were first determined using batch culture, followed by an in-situ experiment deployed at different depths (0.5, 1.5, 3.5, and 5.0 m) in a drinking source water reservoir (Miyun Reservoir, China) to verify the laboratory results. The optimum conditions for growth (7.5 × 108 cells L−1) and MIB production ((1300 ± 29) μg L−1) of Planktothrix sp. were achieved at 85 μmol photons m−2 s−1 in the laboratory and at 1.5 m (the corresponding average light intensity of 66 μmol photons m−2 s−1) in the field. The minimum light requirement for the growth of Planktothrix sp. (4.4 μmol photons m−2 s−1) was determined according to the laboratory data. While the in-situ experiment further indicated that Planktothrix sp. could not successfully grow at depths of 5 m where light intensity was below the minimum light requirement. In addition, the history data also verified the negative relationships between underwater light availability and MIB concentration.  相似文献   

11.
The collision rates between spin-labelled valeric acid in water, and between the corresponding mixed-chain, spin-labelled phosphatidylcholine in water-methanol mixtures, and also between spin-labelled phosphatidylcholine monomers and micelles in water have been determined from the spin-spin broadening of the electron spin resonance spectrum. In each case the second order rate constants are consistent with a diffusion-controlled process. For spin-labelled valeric acid in water the translational diffusion coefficient at 20°C is 3.4 · 10−6 cm2 · s−1, and for spin-labelled phosphatidylcholine varies between 2.3 · 10−6 and 3.8 · 10−6 cm2 · s−1 within the range 44 to 88 wt% methanol. The spin-labelled phosphatidylcholine monomer diffusion coefficient in water at 20°C is 2.4 · 10−6 cm2 · s−1, deduced from the monomer-micelle association rate, with an activation energy of 4.0 kcal · mol−1. The much slower on-rates for association of lipid monomers with phospholipid bilayer vesicles reported in the literature, therefore indicate that incorporation into bilayers is not a diffusion-controlled process.  相似文献   

12.
Autotrophic respiration involves the use of fixed carbon by plants for their own metabolism, resulting in the release of carbon dioxide as a by‐product. Little is known of how autotrophic respiration components vary across environmental gradients, particularly in tropical ecosystems. Here, we present stem CO2 efflux data measured across an elevation transect spanning ca. 2800 m in the Peruvian Amazon and Andes. Forest plots from five elevations were studied: 194, 210, 1000, 1500, and 3025 m asl Stem CO2 efflux (Rs) values from each plot were extrapolated to the 1‐ha plot level. Mean Rs per unit stem surface area declined significantly with elevation, from 1.14±0.12 at 210 m elevation to 0.62±0.09 μmol C m−2 s−1 at 3025 m elevation. When adjusted for changing forest structure with elevation, this is equivalent to 6.45±1.12 Mg C ha−1 yr−1 at 210 m elevation to 2.94±0.19 Mg C ha−1 yr−1 at 3025 m elevation. We attempted to partition stem respiration into growth and maintenance respiration components for each site. Both growth and maintenance respiration rates per unit stem showed similar, moderately significant absolute declines with elevation, but the proportional decline in growth respiration rates was much greater. Stem area index (SAI) showed little trend along the transect, with declining tree stature at higher elevations being offset by an increased number of small trees. This trend in SAI is sensitive to changes in forest stature or size structure. In the context of rapid regional warming over the 21st century, such indirect, ecosystem‐level temperature responses are likely to be as important as the direct effects of temperature on maintenance respiration rates.  相似文献   

13.
In the presence of MgCl2 and ATP, the specific viscosity of suspensions of unsealed freezethawed erythrocyte membranes decreased slowly with time at 37 °C. The decrease in viscosity was found to be an index of Mg-ATP-specific induced folding of these membranes. Mg-ATP-dependent shape or viscosity changes were found to be highly temperature dependent and the viscosity of these membranes did not decrease in the presence of 2 mm 5′-adenyl imidodiphosphate and MgCl2. Cyclic AMP, NaCl, or KCl did not have any effect on the rate of Mg-ATP-induced viscosity decreases. The Mg-ATP-dependent viscosity decreases were inhibited 100% by 1 mm chlorpromazine or 1 mmN-ethylmaleimide. Mg-ATP-dependent viscosity decreases were half-maximally inhibited by 1 μm Ca2+ and completely inhibited by 3–5 μm Ca2+. Ca2+ (5 μm) also inhibited Mg2+-dependent phosphorylation 25 to 30% in these membranes. However, if these membranes were preincubated in the absence of Ca2+ for greater than 10 min at 37 °C, 5 μm Ca2+ no longer inhibited Mg-ATP-dependent viscosity decreases and only inhibited Mg2+-dependent phosphorylation 5% in these preincubated membranes. Preincubation of these membranes at 37 °C for 10 min in the absence of Ca2+ also resulted in the loss of approximately 40 to 50% of the high-Ca2+ affinity Ca + Mg-ATPase activity. The presence of 5 μm Ca2+ in the preincubation medium protected against the loss of the inhibitory effect of Ca2+ on Mg2+-dependent phosphorylation and Mg-ATP-dependent viscosity decreases. The presence of Ca2+ in the preincubation medium also protected against the loss of Ca + Mg-ATPase activity in these membranes. It is hypothesized that freeze-thawed erythrocyte membranes contain a Ca2+ phosphatase activity which is temperature labile in the absence of Ca2+ and that this Ca2+ phosphatase activity may be involved in the regulation of shape of these membranes. Also discussed is the possible relationship of this Ca2+ phosphatase with Ca + Mg-ATPase activity and the problems inherent in studying Ca2+-regulated functions in freeze-thawed erythrocyte membranes.  相似文献   

14.
《Plant science》1987,49(2):75-79
The photosynthetic activity of leaf slices from Spinacia oleracea L., Cucumis sativus L. and Nerium oleander L. was measured in 25° C immediately after preincubation for 2.5 h at various photon flux densities (PFD) with chilling at 4°C, or at a moderate (450 μmol m−2 s−1) PFD with various temperatures below 25°C. Inhibition of photosynthesis was evident in C. sativus and 45°C-grown N. oleander after preincubation at 4°C at all PFD. The inhibition was most severe at fluxes in excess of the moderate PFD under which the plants were grown. Photosynthesis in S. oleracea and 20°C-grown N. oleander was not inhibited at 4°C unless the PFD was in excess of this moderate PFD. The inhibition of photosynthesis was initiated in C. sativus below 13°C, and in 45°C-grown N. oleander below 8°C. A phase transition in the polar lipids from the thylakoids of these plants was detected at about the same temperatures. For S. oleracea and 20°C-grown N. oleander preincubated under the same conditions, there was no inhibition of photosynthesis and no phase transition above 0°C. These results show that some component of photosynthesis was disrupted in the light at temperatures below that of the phase transition in the thylakoid polar lipids.  相似文献   

15.
3-isopropylmalate dehydrogenase (IPMDH) from Escherichia coli was overexpressed, purified and crystallized. The enzyme was characterized and compared to its thermophilic counterpart from Thermus thermophilus strain HB8. As in the thermophile enzyme, the activity of E. coli IPMDH was dependent on the divalent cations, Mg2+ or Mn2+, with Mn2+ being the preferred cation. Activity was also strongly influenced by KCl: 0.3 M were necessary for the optimal activity. At 40°C the Km of E. coli IPMDH was 105 μM for IPM and 321 μM for NAD, the kcat was 69 s−1. The half denaturationn temperature was 64°C, which was 20°C lower than that of the thermophile enzyme.  相似文献   

16.
Water temperature and flow velocity directly affect the fish swimming capacity, and thus, both variables influence the fish passage through river barriers. Nonetheless, their effects are usually disregarded in fishway engineering and management. This study aims to evaluate the volitional swimming capacity of the northern straight-mouth nase (Pseudochondrostoma duriense), considering the possible effects of water temperature, flow velocity and body size. For this, the maximum distance, swim speed and fatigue time (FT) were studied in an outdoor open-channel flume in the Duero River (Burgos, Spain) against three nominal velocities (1.5, 2.5 and 3 m s−1) and temperatures (5.5, 13.5 and 18.5°C), also including the changes between swimming modes (prolonged and sprint). Results showed that a nase of 20.8 cm mean fork length can develop a median swim speed that exceeds 20.7 BL s−1 (4.31 m s−1) during a median time of 3.4 s in sprint mode, or 12.2 BL s−1 (2.55 m s−1) for 23.7 s in prolonged mode under the warmest scenario. During prolonged swimming mode, fish were able to reach further distances in warmer water conditions for all situations, due to a greater swimming speed and FT, whereas during sprint mode, warmer conditions increased the swim speed maintaining the FT. In conclusion, the studied temperature range and flow velocity range influence fish swimming performance, endurance and distance travelled, although with some differences depending on the swimming mode. The provided information goes a step forward in the definition of real fish swimming capacities, and in turn, will contribute to establish clear passage criteria for thermo-velocity barriers, allowing the calculation of the proportion of fish able to pass a barrier under different working scenarios, as well designing of the optimized solutions to improve the fish passage through river barriers.  相似文献   

17.
Bdellovibrio bacteriovorus HD100 is a highly motile predatory bacterium that consumes other Gram-negative bacteria for its sustenance. Here, we describe the impacts the media viscosity has both on the motility of predator and its attack rates. Experiments performed in polyethylene glycol (PEG) solutions, a linear polymer, found a viscosity of 10 mPa s (5% PEG) negatively impacted predation over a 24-h period. When the viscosity was increased to 27 mPa s (10% PEG), predation was nearly abolished. Tests with three other B. bacteriovorus strains, i.e., 109J and two natural isolates, found identical results. Short-term (2-h) experiments, however, found attack rates were improved in 1% PEG, which had a viscosity of 5.4 mPa s, using bioluminescent prey and their viabilities. In contrast, when experiments were performed in dextran, a branched polymer, no increase in predation was seen even though the viscosity was a comparable 5.1 mPa s. The enhanced attack rates in this solution coincided with a 31% increase in B. bacteriovorus HD100 swimming speeds (62 μm s−1 in 1% PEG vs. 47.5 μm s−1 in HEPES-salt).  相似文献   

18.
A chemically defined ram semen diluent (RSD-1) has been developed. RSD-1 maintained spermatozoal motility of diluted semen containing approximately 800 million spermatozoa ml−1 during cooling to 15°C and its storage for 1 h. Motility was further maintained when the cooled semen was diluted to 100 million spermatozoa ml−1 and incubated at 38°C for about 24 h. In contrast, a conventional milk-based diluent supported motility for less than 6 h at 38°C. Spermatozoal motility was influenced by the buffering capacity, osmolarity and the presence or absence of macromolecules and calcium in the chemically defined diluent. Among the organic buffers tested, MOPS (3-(N-morpholino)propanesulphonic acid) had a marked influence on the maintenance of spermatozoal motility. The presence of MOPS also overcame the detrimental effects of 2 mM calcium in Krebs Ringer improved (KR-I) buffer.  相似文献   

19.
β-Mannanase is the key enzyme in the hydrolysis of mannan which has been widely applied in diverse industrial fields such as biobleaching pulps, food and feed industry, bioethanol and pharmaceutical applications. In this study, a novel GH5 family β-mannanase gene (LrMan5B) with 381 amino acid residues was identified from Lichtheimia ramosa, and highly expressed in Pichia pastoris X33. The amino acid sequence shares the highest identity (64%) with the β-mannanase from Rhizomucor miehei. Purified recombinant LrMan5B showed the optimal activity at pH 5.0 and 65 °C. It had broad-range pH stability (retaining >65% activity after incubation at pH 3.0–8.0 at 37 °C for 24 h) and was highly thermostable (retaining >80% activity after incubation at 60 °C for 30 min). LrMan5B displayed the highest catalytic efficiency for locust bean gum and the kcat/Km value was 1357.47 mL·mg−1·s−1, followed by guar gum (512.82 mL·mg−1·s−1), konjac glucomannan (454.21 mL·mg−1·s−1), and palm kernel meal (137.00 mL·mg−1·s−1). In order to evaluate the synergistic effect of LrMan5B and α-galactosidase LrAgal36A from L. ramosa, LrAgal36A was supplemented to hydrolyze palm kernel meal with LrMan5B together, showing that the reducing sugar release significantly increased by 21% (compared with the sum of that by hydrolysis of single Lrman5B or LrAgal36A). Due to its favorable enzymatic properties, LrMan5B might own potential applications in the area of food and feed processing.  相似文献   

20.
The kinetics and mechanism of the oxidation of L- ascorbic acid by trisoxalatocobaltate(III) were studied as a function of pH, ascorbate concentration, ionic strength and temperature in a weakly basic aqueous solution. The pH dependence of the process can be ascribed to the oxidation of the doubly deprotonated ascorbate ion for which k = 20 M−1 s−1 at 25 °C, ΔH# = 34 ± 2 kJ mol−1 and ΔS# = −108 ± 7 J K−1 mol−1. The results are discussed in reference to literature data for this reaction in weakly acidic medium and for the oxidation by a series of other oxidants.  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