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.
An immobilized biocatalyst with invertase activity prepared by immobilization of whole yeast cells without use of any insoluble carrier was tested in tubular fixed-bed reactors from the point of view of possible application for continuous full-scale sucrose hydrolysis. At inlet sucrose concentration above 60% (w/w) and reaction temperature 60–70°C, total sucrose hydrolysis was achieved at a flow rate of 0.6–1.5 bed volumes per hour. At a flow rate about 10 bed volumes per hour, the conversion was still 0.5. The specific productivity of the biocatalyst was 3–25 h−1; the productivity of the reactor was 1–9 kg l−1 h−1. The half-life of the biocatalyst invertase activity was 815 h at 70°C. The specific pressure drop over the biocatalyst bed was less than 23 kPa m−1. The biocatalyst was proved to be fully capable of continuous sucrose hydrolysis in fixed-bed reactors.  相似文献   

2.
The photosynthetic capacity of Myriophyllum salsugineum A.E. Orchard was measured, using plants collected from Lake Wendouree, Ballarat, Victoria and grown subsequently in a glasshouse pond at Griffith, New South Wales. At pH 7.00, under conditions of constant total alkalinity of 1.0 meq dm−3 and saturating photon irradiance, the temperature optimum was found to be 30–35°C with rates of 140 μmol mg−1 chlorophyll a h−1 for oxygen production and 149 μmol mg−1 chlorophyll a h−1 for consumption of CO2. These rates are generally higher than those measured by other workers for the noxious Eurasian water milfoil, Myriophyllum spicatum L., of which Myriophyllum salsugineum is a close relative. The light-compensation point and the photon irradiance required to saturate photosynthetic oxygen production were exponentially dependent on water temperature. Over the temperature range 15–35°C the light-compensation point increased from 2.4 to 16.9 μmol (PAR) m−2 s−1 for oxygen production while saturation photon irradiance increased from 41.5 to 138 μmol (PAR) m−2 s−1 for oxygen production and from 42.0 to 174 μmol (PAR) m−2 s−1 for CO2 consumption. Respiration rates increased from 27.1 to 112.3 μmol (oxygen consumed) g−1 dry weight h−1 as temperature was increased from 15 to 35°C. The optimum temperature for productivity is 30°C.  相似文献   

3.
Comparative measurements of bacterial total counts and volumes of flow cytometry (FCM), transmission electron (TEM), and epifluorescence microscopy (EFM), were undertaken during a four week mesocosm experiment. Total counts of bacteria measured by TEM, EFM, and FCM were in the range of 1 · 106−6 cells ml−1, 1 · 106−3 · 1016 cells ml−1, and 5 · 105 cells ml−1 respectively. The mean volume of the bacterial community, measured by means of EFM and TEM, increased from 0.12–0.15 μm3 at the start of the experiment to 0.39–0.53 μm3 at the end. Generally, there was good agreement between the two methods and regression analyses gave r = 0.87 (p < < 0.01) for cell volume and r = 0.97 (p < < 0.01) for cell number. DAPI stained bacteria with volumes less than 0.2 μm3 were not detected by flow cytometry and these were generally an order of magnitude lower than counts made by TEM and EFM. For samples where the mean bacterial cell volume was longer than 0.3 μm3, all three methods were in agreement both with respect to counts and volume estimates.  相似文献   

4.
Cross-linked waxy maize (CWM) starch dispersions (STDs) of concentration 50 g kg−1 were heated in sucrose solutions containing 0–600 g kg−1 (g sucrose/kg dispersion) at 85 °C at low shear and in intermittently agitated cans at 110 °C. The STDs heated in 0–300 g kg−1 sucrose exhibited antithixotropic behavior, while those heated in 400–600 g kg−1 sucrose exhibited thixotropic behavior. The mean starch granule diameter of the starch dispersions did not show strong dependence on sucrose concentration. The dispersions, especially those with high sucrose concentrations and heated at 110 °C, exhibited G′ versus frequency (ω) profiles of gels. The STDs exhibited first normal stress differences that increased in magnitude with the concentration of sucrose. Values of the first normal stress coefficient of canned dispersions calculated from dynamic rheological data plotted against ω and experimental values plotted against shear rate of some of the STDs overlapped.  相似文献   

5.
A novel nutrient removal/waste heat utilization process was simulated using semicontinuous cultures of the thermophilic cyanobacterium Fischerella. Dissolved inorganic carbon (DIC)-enriched cultures, maintained with 10 mg l−1 daily productivity, diurnally varying temperature (from 55°C to 26–28°C), a 12:12 light cycle (200 μE sec−1 m−2) and 50% biomass recycling into heated effluent at the beginning of each light period, removed > 95% of NO3 + NO2−N, 71% of NH3-N, 82% of PO43− −P, and 70% of total P from effluent water samples containing approximately 400 μg l−1 combined N and 60 μg l−1 P. Nutrient removal was not severely impaired by an altered temperature gradient, doubled light intensity, or DIC limitation. Recycling 75% of the biomass at the end of each light period resulted in unimpaired NO3 + NO2 removal, 38–45% P removal and no net NH3 removal. Diurnally varying P removal, averaging 50–60%, and nearly constant > 80% N removal, are therefore projected for a full-scale process with continuous biomass recycling.  相似文献   

6.
The growth of the freshwater microalga Scenedesmus obliquus was studied at 30°C in a mineral culture medium with phosphorus concentrations of between 0 and 372 μ . The values for the specific growth rates, between and , fitted a semistructured substrate-limitation model with μm1 = 0·0466 h−1, μm2 = 0·0256 h−1 and . The specific uptake rate of phosphorus reached a maximum value of qSm1 = 658·01 × 10−4 μmol P mg−1 biomass h−1.  相似文献   

7.
σ-Methyl-(η5-indenyl) chromium tricarbonyl (III) rearranges quantitatively into η6-1-endo-methylindene) chromium tricarbonyl (IV) in C6D6 solution at 30–60°C. Methyl group attachment to the positions 2 or 3 of indenyl ligand in (III) has no influence on the activation parameters of this ricochet inter-ring haptotropic rearrangement (ΔG#=23.6 kcal mol−1; ΔH#=18.9±0.2 kcal mol−1; ΔS#=−18.6±0.2 cal K−1 mol−1). (IV) undergoes further irreversible isomerization at 60–120° into (ν6-3-methylindene) chromium tricarbonyl (V) with a higher activation barrier (ΔG#=28.5±0.1 kcal mol−1) via two consecutive [1,5]-sigmatropic hydrogen shifts. The mechanisms of both rearrangements have been studied in detail using density functional theory (DFT) calculations with extended basis sets. Calculations show that the rearrangement (III) → (IV) proceeds in two steps. Methyl group migration from chromium into position 1 of the indenyl ligand is the rate-determining step leading to the formation of the 16-electron intermediate (VII). The calculated activation barrier (Ea=19.6 kcal mol−1) is in good agreement with the experimental one. Further rearrangement (VII) → (V) proceeds via a trimethylenemethane-type transition state (XVIII) with an activation barrier 11.8 kcal mol−1. The coordination of the chromium tricarbonyl group at the six-membered ring has only minor influence on the kinetic parameters of the hydrogen [1,5]-sigmatropic shift in indene.  相似文献   

8.
Five coordinate complex formation of zinc(II)–octaethylporphyrin (Zn(OEP)) and zinc(II)–tetraphenylporphyrin (Zn(TPP)) have been studied in the presence of seven systematically selected electron donor molecules. Stability constants in toluene were determined at various temperature ranged between 20 and 60°C by absorption and steady-state fluorescent measurements from which thermodynamic parameters were determined. Depending on the porphyrin and the axial ligand the entropy is changing between −127 and −61 J mol−1 K−1 while the enthalpy is ranging from −49 to −24 kJ mol−1.  相似文献   

9.
The bioconversion of propionitrile to propionamide was catalysed by nitrile hydratase (NHase) using resting cells of Microbacterium imperiale CBS 498-74 (formerly, Brevibacterium imperiale). This microorganism, cultivated in a shake flask, at 28 °C, presented a specific NHase activity of 34.4 U mgDCW−1 (dry cell weight). The kinetic parameters, Km and Vmax, tested in 50 mM sodium phosphate buffer, pH 7.0, in the propionitrile bioconversion was evaluated in batch reactor at 10 °C and resulted 21.6 mM and 11.04 μmol min−1 mgDCW−1, respectively. The measured apparent activation energy, 25.54 kJ mol−1, indicated a partial control by mass transport, more likely through the cell wall.

UF-membrane reactors were used for kinetic characterisation of the NHase catalysed reaction. The time dependence of enzyme deactivation on reaction temperature (from 5 to 25 °C), on substrate concentrations (from 100 to 800 mM), and on resting cell loading (from 1.5 to 200 μg  ml−1) indicated: lower diffusional control (Ea=37.73 kJ mol−1); and NHase irreversible damage caused by high substrate concentration. Finally, it is noteworthy that in an integral reactor continuously operating for 30 h, at 10 °C, 100% conversion of propionitrile (200 mM) was attained using 200 μg  ml−1 of resting cells, with a maximum volumetric productivity of 0.5 g l−1 h−1.  相似文献   


10.
A thermophilic Bacillus sp. strain AN-7, isolated from a soil in India, produced an extracellular pullulanase upon growth on starch–peptone medium. The enzyme was purified to homogeneity by ammonium sulfate precipitation, anion exchange and gel filtration chromatography. The optimum temperature and pH for activity was 90 °C and 6.0. With half-life time longer than one day at 80 °C the enzyme proves to be thermostable in the pH range 4.5–7.0. The pullulanase from Bacillus strain lost activity rapidly when incubated at temperature higher than 105 °C or at pH lower than 4.5. Pullulanase was completely inhibited by the Hg2+ ions. Ca2+, dithiothreitol, and Mn2+ stimulated the pullulanase activity. Kinetic experiments at 80 °C and pH 6.0 gave Vmax and Km values of 154 U mg−1 and 1.3 mg ml−1. The products of pullulan were maltotriose and maltose. This proved that the purified pullulanase (pullulan-6-glucanohydrolase, EC 3.2.1.41) from Bacillus sp. AN-7 is classified under pullulanase type I. To our knowledge, this Bacillus pullulanase is the most highly thermostable type I pullulanase known to date.  相似文献   

11.
P.Muir Wood 《BBA》1974,357(3):370-379
The rate of electron transfer between reduced cytochrome ƒ and plastocyanin (both purified from parsley) has been measured as k = 3.6 · 107 M−1 · s−1, at 298 °K and pH 7.0, with activation parameters ΔH = 44 kJ · mole−1 and ΔS = +46 J · mole−1 · °K−1. Replacement of cytochrome ƒ with red algal cytochrome c-553, Pseudomonas cytochrome c-551 and mammalian cytochrome c gave rates at least 30 times slower: k = 5 · 105, 7.5 · 105 and 1.0 · 106 M−1 · s−1, respectively.

Similar measurements made with azurin instead of plastocyanin gave k = 6 · 106 and approx. 2 · 107 M−1 · s−1 for reaction of reduced azurin with cytochrome ƒ and algal cytochrome respectively.

Rate constants of 115 and 80 M−1 · s−1 were found for reduction of plastocyanin by ascorbate and hydroquinone at 298 °K and pH 7.0. The rate constants for the oxidation of plastocyanin, cytochrome ƒ, Pseudomonas cytochrome c-551 and red algal cytochrome c-553 by ferricyanide were found to be between 3 · 104 and 8 · 104 M−1 · s−1.

The results are discussed in relation to photosynthetic electron transport.  相似文献   


12.
J. Butler  G.G. Jayson  A.J. Swallow 《BBA》1975,408(3):215-222

1. 1. The superoxide anion radical (O2) reacts with ferricytochrome c to form ferrocytochrome c. No intermediate complexes are observable. No reaction could be detected between O2 and ferrocytochrome c.

2. 2. At 20 °C the rate constant for the reaction at pH 4.7 to 6.7 is 1.4 · 106 M−1 · s−1 and as the pH increases above 6.7 the rate constant steadily decreases. The dependence on pH is the same for tuna heart and horse heart cytochrome c. No reaction could be demonstrated between O2 and the form of cytochrome c which exists above pH ≈ 9.2. The dependence of the rate constant on pH can be explained if cytochrome c has pKs of 7.45 and 9.2, and O2 reacts with the form present below pH 7.45 with k = 1.4 · 106 M−1 · s−1, the form above pH 7.45 with k = 3.0 · 105 M−1 · s−1, and the form present above pH 9.2 with k = 0.

3. 3. The reaction has an activation energy of 20 kJ mol−1 and an enthalpy of activation at 25 °C of 18 kJ mol−1 both above and below pH 7.45. It is suggested that O2 may reduce cytochrome c through a track composed of aromatic amino acids, and that little protein rearrangement is required for the formation of the activated complex.

4. 4. No reduction of ferricytochrome c by HO2 radicals could be demonstrated at pH 1.2–6.2 but at pH 5.3, HO2 radicals oxidize ferrocytochrome c with a rate constant of about 5 · 105–5 · 106 M−1 · s−1

.  相似文献   


13.
Arterial pH, PCO2 (PaCO2), plasma bicarbonate [HCO3 and respiratory frequency were measured in pigeons exposed to ambient temperatures (TaS) of 30–60°C. Acclimated, nonpanting birds regulated acid-base balance at normal levels, when exposed to Tas) between 30 and 53°C Ta. At higher Tas (55–60°C), both nonpanting and panting acclimated pigeons regulated pH at normal levels, 7.544 ± 0.011 (SD) and 7.531 ± 0.022 (SD), respectively, accompanied by a slight hypocapnia, 24.8 ± 4.0 Torr and 23.8 ± 2.49 Torr (PaCO2), respectively. Nonacclimated birds, exposed to 50°C Ta, endured a severe hypocapnia (PaCO2 of 9.1 ± 2.52 Torr) and alkalosis (pH of 7.702 ± 0.048). Thirteen exposures to > 50°C Ta, 4–6 h a day, resulted in a significant improvement in the capacity of the panting pigeon to maintain an almost normal acid-base balance, i.e. actual and standard [HCO3 of 22.6 ± 1.22 and 25.7 ± 1.10 mM/l, respectively, and only a slight hypocapnia (PaCO2 of 23.6 ± 3.9 Torr) and alkalosis (pH of 7.589). The suggestion that acclimation to high Tas (50–60°C) is needed for fine adjustment between the competing needs for heat dissipation, pulmonary gas exchange, and acid-base regulation in the heat-exposed pigeon is discussed.  相似文献   

14.
High-pressure liquid-chromatography and microcalorimetry have been used to determine equilibrium constants and enthalpies of reaction for the disproportionation reaction of adenosine 5′-diphosphate (ADP) to adenosine 5′-triphosphate (ATP) andadenosine 5′-monophosphate (AMP). Adenylate kinase was used to catalyze this reaction. The measurements were carried out over the temperature range 286 to 311 K, at ionic strengths varying from 0.06 to 0.33 mol kg−1, over the pH range 6.04 to 8.87, and over the pMg range 2.22 to 7.16, where pMg = -log a(Mg2+). The equilibrium model developed by Goldberg and Tewari (see the previous paper in this issue) was used for the analysis of the measurements. Thus, for the reference reaction: 2 ADp3− (ao) AMp2− (ao)+ ATp (ao), K° = 0.225 ± 0.010, ΔG° = 3.70 +- 0.11 kJ mol −1, ΔH° = −1.5 ± 1. 5 kJ mol −1, °S ° = −17 ± 5 J mol−1 K−1, and ACPp°≈ = −46 J mo1l−1 K−1 at 298.15 K and 0.1 MPa. These results and the thermodynamic parameters for the auxiliary equilibria in solution have been used to model the thermodynamics of the disproportionation reaction over a wide range of temperature, pH, ionic strength, and magnesium ion morality. Under approximately physiological conditions (311.15 K, pH 6.94, [Mg2+] = 1.35 × 10−3 mol kg−1, and I = 0.23 mol kg−1) the apparent equilibrium constant (KA′ = m(ΣAMP)m(ΣATP)/[ m(ΣADP)]2) for the overall disproportionation reaction is equal to 0.93 ± 0.02. Thermodynamic data on the disproportionation reaction and literature values for this apparent equilibrium constant in human red blood cells are used to calculate a morality of 1.94 × 10−4 mol kg−1 for free magnesium ion in human red blood cells. The results are also discussed in relation to thermochemical cycles and compared with data on the hydrolysis of the guanosine phosphates.  相似文献   

15.
The infrared, visible and nuclear magnetic resonance spectra of protochlorophyll a and vinylprotochlorophyll a in dry non-polar solvents (carbon tetrachloride, chloroform, cyclohexane) are presented and interpreted in terms of dimer interaction.

The infrared spectra in the 1600–1800 cm−1 region clearly show the existence of a coordination interaction between the C-9 ketone oxygen function of one molecule and the central magnesium atom of another molecule. Infrared spectra in the OH stretching region (3200–3800 cm−1) provide a valuable test of the water content in the samples.

The analysis of the absorption and circular dichroism spectra of protochlorophyll a and vinylprotochlorophyll a in carbon tetrachloride demonstrates the existence of a monomer-dimer equilibrium in the concentration range from 10−6 to 5 · 10−4 M. The dimerization constants are (6±2) · 105 1 · M−1 for protochlorophyll a and (4.5±2) · 105 1 · M−1 for vinylprotochlorophyll a at 20 °C. The deconvolution of visible spectra in the red region has been performed in order to obtain quantitative information on the dimer structure. Two models involving a parallel or a perpendicular arrangement of the associated molecules are considered.

From 1H NMR spectra, it appears that the region of overlap occurs near ring V, in agreement with the interpretation of the infrared spectra.  相似文献   


16.
1. Skin and rectal temperatures were recorded continuously in 70 measurements during typical tasks of infantry and artillery training at 0 to −29 °C. The duration of the measurements varied from 55 min to 9.5 h.

2. The distribution of finger skin temperatures was quite similar at ambient temperature ranges 0 to −10 °C and −10 to −20 °C, while at −20 to −30 °C the finger temperatures were clearly lower.

3. At different ambient temperature ranges, 20–69% of finger temperatures were low enough to cause cold thermal sensations.

4. Sensation of cold was experienced at a finger temperature of 11.6±3.7 °C (mean±SD).  相似文献   


17.
Seagrasses are recognized as important plant communities in coastal estuaries and lagoons across both tropical and temperate climes; thus, large-scale seagrass die-off events worldwide are of general concern. In Florida Bay, at the southern terminus of the Florida peninsula, seagrass die-off events up to 4000 ha have been reported and smaller scale mortality events are noted annually. In the present study, we examined several hypothesized causative factors (high temperature, hypersalinity, sulfide toxicity) of seagrass (Thalassia testudinum) mortality in Florida Bay. To test sulfide effects, in situ sulfide production was stimulated by applying a labile carbon source (glucose) to sulfate reducers in the sediment at five sites across the bay (northeastern, northcentral, and southwestern basins). During the one year study, high temperature (32–36 °C) and salinity (> 50 psu) were recorded in the bay associated with a regional drought. We also experienced major seagrass die-off events at two of our southwestern bay sites. These field conditions provided an excellent opportunity to closely examine cause–effect relationships among stressors and die-off events in the field, and verify results of our previous mesocosm experiments. Even though glucose amendments stimulated porewater sulfides in bay sediments (4–8 mmol L− 1), no significant differences in biomass, short shoot density or final growth rates were found between control and glucose plots. In addition, the highest growth rates and shoot densities were concomitant with maximum water column salinity (> 50 psu) and temperature (32–36 °C), when porewater sulfides were also in the millimolar range. Large-scale seagrass mortality events, encompassing  50% of the entire meadow at one site, occurred at southwestern bay sites when plants were down regulating (slower growth and shoot density), probably in response to shorter day length and lower temperature (30–34 to 23–26 °C) from October, 2004 to January, 2005. Sulfate reduction rates (SRR) were also 2-fold higher in the southwestern (214–488 nmol cm− 3 d− 1) versus northcentral and northeastern (97–240 nmol cm− 3 d− 1) bay sites, possibly limited by labile carbon, which we found to stimulate SRR 3-fold in northeastern and northcentral bay sites (461–708 nmol cm− 3 d− 1) and 4-fold at southwestern bay sites (1211–2036 nmol cm− 3 d− 1). Based on a synthesis of the field data reported herein, our mesocosm experiments to date, and contributions by others, we present a conceptual model of seagrass die-off in Florida Bay outlining a cascade of stressors, stimulated by P enrichment, which leads to high O2 consumption in the system triggering a seagrass die-off event.  相似文献   

18.
We measured the toxicity and mutagenicity induced in human diploid lymphoblasts by various radiation doses of X-rays and two internal emitters. [125I]iododeoxyuridine ([125I]dUrd) and [3H]thymidine ([3H]TdR), incorporated into cellular DNA. [125I]dUrd was more effective than [3H]TdR at killing cells and producing mutations to 6-thioguanine resistance (6TGR). No ouabain-resistant mutants were induced by any of these agents. Expressing dose as total disintegrations per cell (dpc), the D0 for cell killing for [125I]dUrd was 28 dpc and for [3H]TdR was 385 dpc. The D0 for X-rays was 48 rad at 37°C. The slopes of the mutation curves were approximately 75 × 10−8 6TGR mutants per cell per disintegration for [125I]dUrd and 2 × 10−8 for [3H]TdR. X-Rays induced 8 × 10−8 6TGR mutants per cell per rad. Normalizing for survival, [125I]dUrd remained much more mutagenic at low doses (high survival levels) than the other two agents. Treatment of the cells at either 37°C or while frozen at −70°C yielded no difference in cytotoxicity or mutation for [125I]dUrd or [3H]TdR, whereas X-rays were 6 times less effective in killing cells at −70°C.

Assuming that incorporation was random throughout the genome, the mutagenic efficiencies of the radionuclides could be calculated by dividing the mutation rate by the level of incorporation. If the effective target size of the 6TGR locus is 1000–3000 base pairs, then the mutagenic efficiency of [125I]dUrd is 1.0–3.0 and of [3H]TdR is 0.02–0.06 total genomic mutations per cell per disintegration. 125I disintegrations are known to produce localized DNA double-strand breaks. If these breaks are potentially lethal lesions, they must be repaired, since the mean lethal dose (D0) was 28 dpc. The observations that a single dpc has a high probability of producing a mutation (mutagenic efficiency 1.0–3.0) would suggest, however, that this repair is extremely error-prone. If the breaks need not be repaired to permit survival, then lethal lesions are a subset of or are completely different from mutagenic lesions.  相似文献   


19.
In this paper a number of experiments with the purple bacteria Rhodospirillum rubrum and Rhodopseudomonas capsulata is described in which the total fluorescence yield and/or the total fraction of reaction centers closed after a picosecond laser pulse were measured as a function of the pulse intensity. The conditions were such that the reaction centers were either all in the open or all in the closed state before the pulse arrived. These experiments are analysed using the theoretical formalism discussed in the preceding paper (Den Hollander, W.T.F., Bakker J.G.C., and Van Grondelle, R., Biochim. Biophys. Acta 725, 492–507). From the experimental results the number of connected photosynthetic units, λ, the rate of energy transfer between neighboring antenna molecules, kh, and the rate of trapping by an open reaction center, kot, can be estimated. For R. rubrum it is found that λ = 14−17, kh = (1−2)·1012 s−1 and kot = (4−6)·1011 s−1, for Rps. capsulata λ ≈ 30, kh ≈ 4·1011 s−1 and kot ≈ 3·1011 s−1. The findings are discussed in terms of current models for the structure of the antenna and the kinetic properties of the decay processes occurring in these purple bacteria.  相似文献   

20.
NAD+ kinase (ATP: NAD+ 2-phosphotransferase, EC2.7.1.23) isolated from chicken liver was immobilized on a silica-based support possessing aldehyde functional groups. The highest catalytic activity achieved was 16 U g−1 solid. The optimal pH for the catalytic activity of the immobilized NAD+ kinase was pH 7.1–7.3. The apparent optimum temperature for the immobilized enzyme was about 5°C higher than that of the soluble enzyme. There were no significant differences in the Km app values. The immobilization improved the conformational stability of the enzyme. In preliminary experiments, a 95% conversion of NAD+ to NADP+ was achieved with use of the immobilized NAD+ kinase, which preserved its starting activity practically unchanged up to 36 days.  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