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.
The rates of shrinkage at constant temperature, and growth under a temperature rise below 100°C, of bubbles entrained in wheat flour dough were analyzed and compared with those of a bubble in water. The rate of shrinkage of bubbles in flour dough was controlled by the diffusion of dissolved air from the surface of bubbles to the bulk of flour dough. The apparent diffusion coefficient of the dissolved air in wheat flour dough with the water fraction of 0.49 calculated from the shrinkage of bubbles, was (3.2 ± 1.5) × 101?1 m2/sec (19°C), and (6.4 ± 2.0) × 10?11 m2/sec (42°C). However, the growth behavior of bubbles in flour dough under a temperature rise was very different from that predicted from the diffusion theory. The critical radius of bubbles to grow was larger than that estimated from the diffusion theory. The mechanism of growth of bubbles in wheat flour dough, which was different from that of a bubble in water, is a subject that needs to be clarified.  相似文献   

2.
Abstract— Tryptophan transport across the blood-brain barrier was studied using a single injection dual isotope label technique, in the following three conditions: normal rats, rats with portacaval shunts, and rats with portacaval shunts followed 65 h later by hepatic artery ligation. In both normal rats and those with acute hepatic failure the tryptophan transport system was found to be comprised of two kinetically distinct components. One component was saturable and obeyed Michaelis-Menten kinetics (normal: Vmax= 19.5 nmol.min?1.g?1. Km= 113 μM; hepatic failure: Vmax, = 33.8 nmol.min?1.g?1, Km= 108 μM), and the second was a high capacity system which transported tryptophan in direct proportion to concentration over the range tested (normal: K= 0.026 ml.min?1.g?1; hepatic failure: K= 0.067 ml.min?1.g?1). Since the saturable low capacity component transports several neutral amino acids, and their collective plasma concentration is high in relation to the individual Kms, tryptophan transport by this component is reduced by competitive inhibition under physiological conditions. Thus it was calculated that in normal rats approx 40% of tryptophan influx occurs via the high capacity system. During acute hepatic failure transport via both components was increased substantially, approximately doubling the rate of tryptophan penetration of the blood-brain barrier at all concentrations tested. The contribution by the high capacity component became even more significant than in normal rats, accounting for about 75% of all tryptophan passage from plasma to brain. Brain tryptophan content was 29.9 nmol/g in normal rats and rose to 45.2 nmol/g in rats with portacaval shunts and 50.5 nmol/g in those with acute hepatic failure, correlating with the increased rate of tryptophan transport. In a previous study we found that plasma competing amino acids were greatly increased during acute hepatic failure. Calculations predict that these increased concentrations would cause a reduction in tryptophan transport by the low capacity system. However, because of the increase in the rate of transport by the high capacity component, net tryptophan entry across the blood-brain barrier was actually increased. This increased rate of transport clearly contributes to the increased content of brain tryptophan found during hepatic failure.  相似文献   

3.
A sensitive, rapid, and precise method is described for the continuous determination of oxygen in gases. The principle of the method is based on the reaction of O2 with an alkaline catechol or pyrogallol solution, which is combined with a Fe2+ solution to increase the sensitivity of the color reaction. The development of the color takes place in a tube system provided with a proportioning pump and is read automatically on a recorder after passing a flow cell of a photometer. The lower limit of sensitivity of this method is 0.05 μl of O2 min?1. Thus, in a gas flow of ≈0.5 ml min?1, an oxygen concentration of ≥0.01% (v/v) can be determined. If the gas flow is increased to ≈2.5 ml min?1, this limit of sensitivity is lowered to ≥0.002% (v/v). Since a 2-min period is necessary for the measurement, the volume of the sample has to be 1 ml in the first case and 5 ml in the second.  相似文献   

4.
J Greve  J Blok 《Biopolymers》1975,14(1):139-154
Measurements of electric birefringence, sedimentation velocity, and biological adsorption rate are used to study the properties of bacteriophage T4B in the presence of excess tryptophan. The adsorption rate determined in borate buffer pH 9 (at 25°C) increases from 0.003 × 10?8 ml min?1 (0.025 M) to 0.130 × 10?8 ml min?1 (0.150 M). The Kerr coefficient, rotational diffusion coefficient, and the sedimentation coefficient of the phage are also dependent on buffer concentration and reach plateau values above 0.12 M given by Ksp = ?(275 ± 18) × 10?9 OD?1 cm2 statvolt?2, D25,w = 133 ± 4 sec?1, and s20,w = 818 ± 11 S. From a comparison of electric birefringence measurements of T4B and T4D it is concluded that T4D and T4B (in the presence of excess tryptophan) exhibit a similar hydrodynamic behavior. The change in physical parameters is solely due to a shift in fiber configuration. At high buffer concentrations the fibers make an angle of approximately 3π/4 with the sheath and the permanent dipole moment is about 200,000 D. This dipole moment is roughly ten times as large as that of a phage particle with nonextended fibers. This difference may be due to a change in hydrodynamic center upon fiber extension or to the presence of positive charges on the fiber tips, or both. At intermediate buffer concentrations the phage population behaves as if it were monodisperse. Probably not all six fibers are extended under such conditions.  相似文献   

5.
Photosynthesis and respiration rates of blades from a selected, fast growing strain of the marine red alga. Gigartina exasperata Harvey and Bailey, a carrageenan producer, were measured with an oxygen electrode and compared with rates similarly obtained from wild material of the same species. The measurements, expressed as μl O2 · mg chl a?1, min?1. were made over a light intensity range from 5 to 800 μE · m?2 · sec?1 and a temperature range of 6 to 16°C. The photosynthesis light intensity data are best described by hyperbolic functions.  相似文献   

6.
The effects of biofilms on chemical processes in surficial sediments   总被引:7,自引:0,他引:7  
1. The objectives of the present work were: (a) to evaluate the effects of the development and the presence of a photosynthetically active algal biofilm on chemical fluxes and processes at the sediment–water interface; (b) to measure the effects of the biofilm on chemical concentration gradients in the bulk sediment; and (c) to monitor pH and dissolved oxygen concentration in the biofilm, and through the sediment–water interface using microelectrodes. 2. Two experiments were performed over a period of 8 weeks using a recirculating fluvarium channel containing river sediments with an exposed surface of 0.2 m2 and 20 dm3 of overlying solution. The first experiment was in darkness with minimal effects of a photosynthetically active biofilm. The second experiment in natural light produced a complex photosynthetic biofilm involving a succession of diatoms, green algae and cyanobacteria. 3. The solution overlying the biofilm was monitored continuously for dissolved calcium, silicon, phosphorus, alkalinity and oxygen, as well as conductivity, temperature and pH. The surface of the sediment was also monitored for biological and physical changes as the biofilm developed. The overlying solution was analysed over a period of 48 h at 2-h intervals to examine the effects of a well-developed algal biofilm. At the end of the 48 h, pH and oxygen microelectrodes were used to measure gradients above and through the biofilm, and porewaters were analysed from sediments which had been longitudinally sectioned at a maximum depth resolution of 0.5 mm. 4. Biofilm development had a large influence on the composition of the overlying solution and the development of vertical concentration gradients of solutes in the porewater. Once a diatom community was established, the concentration of dissolved silicon was low (< 40 μm ), with all the silicon diffusing from the underlying sediment being consumed in the biofilm (≈ 0.026 μmol m?2 s?1 at the end of the experiment). 5. The concentrations of calcium, alkalinity and phosphorus in digested sediment increased near the sediment–water interface. X-ray diffraction analysis showed that calcite was formed at the surface. Estimation of the fluxes of alkalinity and calcium in the overlying solution were consistent with calcite formation during daylight and possible dissolution in darkness. The maximum precipitation flux of calcium was 0.87 μmol m?2 s?1 and the maximum net release flux was 0.89 μmol m?2 s?1. 6. The net loss of soluble reactive phosphorus from the overlying solution measured over the intensive sampling period of 48 h is consistent with a coprecipitation mechanism and a surface density of phosphorus included in the lattice of calcite of 0.097 μmol m?2. Processes in the biofilm rather than diffusion from the sediment porewater control chemical fluxes of calcium, alkalinity and phosphorus from the sediment to the overlying water.  相似文献   

7.
Three rice varieties, cv. Norin 36, cv. Norin 37 and cv. Yubae, were grown in a loam with a 20 cm water-table which gave aerobic conditions to a depth of not less than 15 to 17 cm. Under these conditions Norin 36 grew more vigorously and tillered more frequently than the other two varieties. The rates of oxygen diffusion at 23°C from roots up to 11 cm in length were however appreciably lower for Norin 36 (4.3 × 10?8g · cm?2 of root surface · min?1) than for Norin 37 or Yubae (c. 7.8 × 10?8g). A considerable increase (up to 200 %) in the oxygen diffusion rate (ODR) from the roots occurred if they were cooled to 3°C, and at this temperature differences in ODR between the varieties were not significant. For a purely physical system, because of the decrease in the diffusion coefficient of oxygen in water, and, the increase in oxygen solubility, a drop of c. 20 % in ODR should accompany the above 20°C drop in temperature. A 16 % drop was recorded for artificial ‘roots’ under these conditions. It was concluded that respiratory activity at the higher temperature must have been responsible for the low readings and intervarietal differences observed at 23°C. By increasing the 3°C values by 25 % a mean value of 14.2 × 10?8g · cm?2 of root surface · min?1 was recorded for the three varieties, being the probable ODR at 23°C in the absence of a respiratory factor. Calculations show that respiratory activity removed enough oxygen to reduce the ODR for Norin 36 by more than 9 × 10?8g, and for Norin 37 and Yubae by c. 6.7 × 10?8g · cm?2 of root surface · min?1. Anatomical investigations showed that cortical breakdown was always extensive at 4 to 4.5 cm from the apex of the roots. In some cases however breakdown had not occurred in the basal segment of the root. No opinion could be formed as to whether differences in the amount of cortical breakdown between the varieties might have occasioned the respiratory differences observed. An interesting feature of the root anatomy was the failure of breakdown in those regions of the roots through which lateral roots emerged.  相似文献   

8.
Uptake of 14C-labelled sucrose and glucose by isolated seed coat halves of pea (Pisum sativum L. cv. Marzia) seeds was measured in the concentration range <0.1 μM to 100 mM. The initial influx of sucrose was strictly proportional to the external concentration, with a coefficient of proportionality (k) of 6.2 μmol·(g FW)?1·min?1·M?1. Sucrose influx was not affected by 10 μM carbonylcyanide m-chlorophenylhydrazone (CCCP), but it was inhibited by 40% in the presence of 2.5 mM p-chloromercuribenzenesulfonic acid (PCMBS). Influx with diffusional kinetics was also observed for glucose (k = 4.8 μmol·(g FW)?1·min ?1·M ?1) and mannitol (k = 5.1 μmol·(g FW)?1·min?1·M?1). For glucose an additional saturable system was found (Km = 0.26 mM, V max = 4.2 nmol·(g FW)?1·min?1), which appeared to be completely inhibited by CCCP and partly by PCMBS. In contrast to the diffusional pathway, uptake by this saturable system was slightly pH-dependent, with an optimum at pH 5.5. The influx of sucrose appears to be by the same pathway as the efflux of endogenous sucrose, which was inhibited by 36% in the presence of 2.5 mM PCMBS (De Jong A, Wolswinkel P, 1995, Physiol Plant 94: 78–86). It is argued that passive transport may be the only mechanism for sucrose transport through the plasma membrane of seed coat parenchyma cells. The estimated permeability coefficient of the plasma membrane for sucrose (P = 3.5·10?7 cm·s?1) is more than 1 × 106-fold higher than that reported for artificial lipid membranes. This relatively high permeability is hypothesized to result from pore-forming proteins that allow the diffusion of sucrose. Furthermore, it is shown that a sucrose gradient across the plasma membrane of the seed coat parenchyma of only 22 mM will suffice to result in the net efflux of sucrose which is required to feed the embryo.  相似文献   

9.
The effect of irradiance and temperature on the photosynthesis of the red alga, Pyropia tenera, was determined for maricultured gametophytes and sporophytes collected from a region that is known as one of the southern limits of its distribution in Japan. Macroscopic gametophytes were examined using both pulse‐amplitude modulated fluorometry and/or dissolved oxygen sensors. A model of the net photosynthesis–irradiance (P‐E) relationship of the gametophytes at 12°C revealed that the net photosynthetic rate quickly increased at irradiances below the estimated saturation irradiance of 46 μmol photons m?2 s?1, and the compensation irradiance was 9 μmol photons m?2 s?1. Gross photosynthesis and dark respiration for the gametophytes were also determined over a range of temperatures (8–34°C), revealing that the gross photosynthetic rates of 46.3 μmol O2 mgchl‐a?1 min?1 was highest at 9.3 (95% Bayesian credible interval (BCI): 2.3–14.5)°C, and the dark respiration rate increased at a rate of 0.93 μmol O2 mgchl‐a?1 min?1°C?1. The measured dark respiration rates ranged from ?0.06 μmol O2 mgchl‐a?1 min?1 at 6°C to ?25.2 μmol O2 mgchl‐a?1 min?1 at 34°C. The highest value of the maximum quantum yield (Fv/Fm) for the gametophytes occurred at 22.4 (BCI: 21.5–23.3) °C and was 0.48 (BCI: 0.475–0.486), although those of the sporophyte occurred at 12.9 (BCI: 7.4–15.1) °C and was 0.52 (BCI: 0.506–0.544). This species may be considered well‐adapted to the current range of seawater temperatures in this region. However, since the gametophytes have such a low temperature requirement, they are most likely close to their tolerable temperatures in the natural environment.  相似文献   

10.
The relationship between leaf resistance to water vapour diffusion and each of the factors leaf water potential, light intensity and leaf temperature was determined for leaves on seedling apple trees (Malus sylvestris Mill. cv. Granny Smith) in the laboratory. Leaf cuticular resistance was also determined and transpiration was measured on attached leaves for a range of conditions. Leaf resistance was shown to be independent of water potential until potential fell below — 19 bars after which leaf resistance increased rapidly. Exposure of leaves to CO2-free air extended the range for which resistance was independent of water potential to — 30 bars. The light requirement for minimum leaf resistance was 10 to 20 W m?2 and at light intensities exceeding these, leaf resistance was unaffected by light intensity. Optimum leaf temperature for minimum diffusion resistance was 23 ± 2°C. The rate of change measured in leaf resistance in leaves given a sudden change in leaf temperature increased as the magnitude of the temperature change increased. For a sudden change of 1°C in leaf temperature, diffusion resistance changed at a rate of 0.01 s cm?1 min?1 whilst for a 9°C leaf temperature change, diffusion resistance changed at a rate of 0.1 s cm?1 min?1. Cuticular resistance of these leaves was 125 s cm?1 which is very high compared with resistances for open stomata of 1.5 to 4 s cm?1 and 30 to 35 s cm?1 for stomata closed in the dark. Transpiration was measured in attached apple leaves enclosed in a leaf chamber and exposed to a range of conditions of leaf temperature and ambient water vapour density. Peak transpiration of approximately 5 × 10?6 g cm?2 s?1 occurred at a vapour density gradient from the leaf to the air of 12 to 14 g m?3 after which transpiration declined due presumably to increased stomatal resistance. Leaves in CO2-free air attained a peak transpiration of 11 × 10?6 g cm?2 s?1 due to lower values of leaf resistance in CO2 free air. Transpiration then declined in these leaves due to development of an internal leaf resistance (of up to 2 s cm?1). The internal resistance was masked in leaves at normal CO2 concentrations by the increase in stomatal resistance.  相似文献   

11.
Organic electronic devices (OEDs), e.g., organic solar cells, degrade quickly in the presence of ambient gases, such as water vapor and oxygen. Thus, in order to extend the lifetime of flexible OEDs, they have to be protected by encapsulation. A solution‐based encapsulation method is developed, which allows the direct deposition of the diffusion barrier on top of OEDs, thus avoiding lamination of barrier films. The method is based on the deposition of a perhydropolysilazane (PHPS) ink and its subsequent conversion into a silica layer by deep UV irradiation. The resulting barrier films show water vapor transmission rates (WVTRs) of <10?2 g m?2 d?1 (40 °C/85% relative humidity (RH)) and oxygen transmission rates (OTRs) of <10?2 cm3 m?2 d?1 bar?1 at ambient conditions. Flexibility of the resulting barrier films is improved by coating a barrier stack of several thin PHPS layers alternating with organic polymer interlayers. These stacks show an increase of WVTR values by less than 10% after 3000 bending cycles. Direct coating of the PHPS films on top of organic solar cells enhances the device lifetime in damp heat conditions from a few hours to beyond 300 h.  相似文献   

12.
Macrophyte combined with artificial aeration is a promising in situ remediation approach for urban rivers polluted with nutrients and organic matter. However, seasonal variations and aeration effects on phytoremediation performance and root-adhered microbial communities are still unclear. In this study, Pontederia cordata was used to treat polluted urban river water under various aeration intensities. Results showed that the highest removal efficiencies of chemical oxygen demand (CODCr) and total nitrogen (TN) were attained under aeration of 30 L min?1 in spring and summer and 15 L min?1 in autumn, while total phosphorus (TP) removal reached maximum with aeration of 15 L min?1 in all seasons. Moderate aeration was beneficial for increasing the diversity of root-adhered bacteria communities, and the shift of bacterial community structure was more pronounced in spring and autumn with varying aeration intensity. The dual effect, i.e. turbulence and dissolved oxygen (DO), of aeration on the removal of CODCr and TN prevailed over the individual effect of DO, while DO was the most influential factor for TP removal and the root-adhered bacterial community diversity. P. cordata combined with 15 L min?1 aeration was deemed to be the best condition tested in this study.  相似文献   

13.
Abstract

Saturation experiments were performed on intact human peripheral mononuclear leucocytes (MNL) and MNL membranes with (-)125Iodocyanopindolol (125ICYP) over a large concentration range (1.5-600pmol/l). The corresponding Scatchard plots were curvilinear suggesting two saturable classes of binding sites: A high affinity binding site (Bmax1=1000±400 sites/cell, Kd1= 2.1±0.9 pmol/l for intact MNL and Bmax1=550±190 sites/cell, Kd1=4.1±0.9 pmol/l for MNL membranes)and a low affinity binding site (Bmax2=9150±3590 binding sites/cell, Kd2=440±50 pmol/l for intact MNL and Bmax2=11560±4690 sites/cell, Kd2=410±70 pmol/l for MNL membranes). Dissociation of (-)125ICYP from MNL was biphasic consisting of a slow dissociating component (dissociation rate constant k-1=(0.5±0.2)x10?3 min?1 for intact MNL and k-1=(1.0±0.1)x10?3min?1 for MNL membranes) and a fast dissociating component (k-2=(80±20)x10?3min?1 for intact MNL and k-2=(60±10)x10?3min?1 for MNL membranes). In dissociation experiments started after equilibration with various (-)125ICYP concentrations k-1 and k-2 were independent of the equilibrium concentration, whereas the percentual occupancy of the slow and the fast dissociating component varied and was similar to the estimated fractional occupancy of either binding site at the same (-)125ICYP concentrations in saturation experiments. The association rate constant was in the same order of magnitude for both binding sites. These results suggest two independent classes of binding sites for (-)125ICYP on MNL.  相似文献   

14.
The aim of this paper was to study the effects of temperature and irradiance on the photodegradation state of killed phytoplankton cells. For this purpose, killed cells of the diatom Chaetoceros neogracilis RCC2022 were irradiated (photosynthetically active radiation) at 36 and 446 J · s?1 · m?2 (for the same cumulative dose of irradiation energy) and at two temperatures (7°C and 17°C). Analyses of specific lipid tracers (fatty acids and sterols) revealed that low temperatures and irradiances increased photooxidative damages of monounsaturated lipids (i.e., palmitoleic acid, cholesterol and campesterol). The high efficiency of type II photosensitized degradation processes was attributed to: (i) the relative preservation of the sensitizer (chlorophyll) at low irradiances allowing a longer production of singlet oxygen and (ii) the slow diffusion rate of singlet oxygen through membranes at low temperatures inducing more damages. Conversely, high temperatures and irradiances induced (i) a rapid degradation of the photosensitizer and a loss of singlet oxygen by diffusion outside the membranes (limiting type II photosensitized oxidation), and (ii) intense autoxidation processes degrading unsaturated cell lipids and oxidation products used as photodegradation tracers. Our results may explain the paradoxical relationship observed in situ between latitude and photodegradation state of phytoplankton cells.  相似文献   

15.
Suspensions of 2 to 5% rat thymocytes were incubated at 35 °C in buffered balanced salt solution (pH 7.3) with lactate and β-hydroxybutyrate as fuels. The dependence of 3-O-[Me-3H]methylglucose influx on external and internal 3-O-methylglucose concentrations was studied. Entry was almost rectilinear during the first minute. From the dependence of methylglucose entry (into sugar-free cells) on external methylglucose concentration, we judged the entry Km to be about 7.7 mm and the entry V to be about 0.64 μmol · min?1 · (ml of packed cell volume)?1. Methylglucose inside the cell enhanced influx, hence equilibrium exchange was faster than entry. The dependence of equilibrium exchange on methylglucose concentration (inside and outside being equal) indicated a Km of about 25 mm and a V of about 2.1 μmol · (min)?1 · (ml of cell volume)?1. This effect of internal sugar indicated that entry into sugar-free cells is limited mainly by the return of empty carrier to the outside surface and that loading the carrier on the inside enhances its outward mobility. The Km and V for influx into cells containing 21 mm methylglucose were 5.9 mm and 1.17 μmol · min?1 · (ml of packed cells)?1. The effect of 21 mm internal sugar on lowering the influx Km from about 7.7 mm to about 6 mm was reproducible and contributed to the evaluation of the constants of the transport rate law. It indicated that loading of the carrier at the external surface reduces its mobility, in contrast to the effect of loading on the inside. Mechanical explanations for this behavior are discussed.  相似文献   

16.
The reactions of 4-chloronitrosobenzene with pyruvate decarboxylase and transketolase were investigated by use of a new high-pressure liquid chromatography method to determine any differences between these two enzymes with respect to hydroxamic acid production. In addition to the previously established difference in the type of hydroxamic acid produced by the two enzymes, several new and interesting differences in their reaction with nitrosoaromatics were discovered. Most notable was the finding that pyruvate decarboxylase gave 4-chlorophenylhydroxylamine as the major product from 4-chloronitrosobenzene, while transketolase did not produce any detectable hydroxylamine. A redox mechanism was proposed to account for arylhydroxylamine production by pyruvate decarboxylase. This redox mechanism can also explain hydroxamic acid production by pyruvate decarboxylase; however, a previously proposed nucleophilic reaction mechanism occurring simultaneously could not be totally disproven. Either of the two mechanisms is equally likely for transktolase action in view of the present evidence. Another major difference between these enzymes is that the rate of 4-chloronitrosobenzene conversion was found to be much faster for pyruvate decarboxylase than for transketolase when each enzyme was subjected to its own optimal reaction conditions. Transketolase displayed typical enzyme saturation kinetics with 4-chloronitrosobenzene with a Km of 0.31 mM and Vmax of 0.033 μmol ml?1 min?1 unit?1 relative to 5 mMd-fructose 6-phosphate as sugar substrate. On the other hand, the reaction with pyruvate decarboxylase was first order in 4-chloronitrosobenzene with a combined rate constant of 2.0 min?1 unit?1 ml.  相似文献   

17.
A two-dimensional model for substrate transfer and biodegradation in a novel, annular fiber-illuminating bioreactor (AFIBR) is proposed in which photosynthetic bacteria are immobilized on the surface of a side-glowing optical fiber to form a stable biofilm. When excited by light, the desired intensity and uniform light distribution can be obtained within the biofilm zone in bioreactor and then realize continuous hydrogen production. Substrate transfer and biodegradation within the biofilm zone, as well as substrate diffusion and convection within bulk fluid regions are considered simultaneously in this model. The validity of the model is verified experimentally. Based on the model analysis, influences of flow rate and light intensity on the substrate consumption rate and substrate degradation efficiency were investigated. The simulation results show that the optimum operational conditions for the substrate degradation within the AFIBR are: flow rate 100 ml h?1 and light intensity 14.6 μmol photons m?2 s?1.  相似文献   

18.
Anacystis nidulans grown under high and low light, 100 and 10 μE m?2 s?1, respectively, was analyzed with respect to chlorophyll/P700, phycobiliproteins/P700, chlorophyll/cell, and oxygen evolution parameters. The photosynthetic unit sizes of this cyanobacterium, measured as the ratio of total chromophores (chlorophyll and bilin) to P700, were shown to be similar to those of higher plants and green algae. High light grown cells possessed a photosynthetic unit consisting of a core of 157 ± 6 chlorophyll a molecules per P700 associated with a light harvesting system of 95 ± 3.5 biliprotein chromophores. Low light grown cells had substantially more biliprotein chromophores per P700 (125 ± 3.1) than high light cells, but showed no significant difference in the numbers of chlorophyll a molecules per P700 (149 ± 4). Analyses of aqueous biliprotein extracts indicate that low light grown cells produce proportionately more phycocyanin relative to allophycocyanin than high light cells. Calculations of the molecular weight of biliproteins per P700 suggest that there is less than one phycobilisome per reaction center I under both growth conditions. Differences in chlorophyll/cell ratios and oxygen evolution characteristics were also observed. High light cells contain 6.3 × 10?12 mg chlorophyll cell?1, while low light grown cells contain 12.8 × 10?12 mg chlorophyll cell?1. Photosynthetic oxygen evolution rate vs. light intensity curves indicate that high light grown cells reach maximal levels of oxygen evolution at higher light intensity than low light grown cells. Maximal rates of oxygen evolution were 16.6 μmol oxygen min?1 (mg chlorophyll)?1 for high and 8.4 μmol oxygen min?1 (mg chlorophyll)?1 for low light cells. Maximal oxygen evolution rates per cell were equivalent for both cell types, although the amount of P700 per cell was lower in high light cells. High light grown cells are therefore capable of producing more oxygen per reaction center I than low light grown cells.  相似文献   

19.
Submitochondrial particles (SMP) were produced from Jerusalem artichoke (Helianthus tuberosus L.) mitochondria by sonication and differential centrifugation. The SMP were about 50% inside-out as measured by the access of reduced cytochrome c to cytochrome c oxidase. Uncoupled NADH oxidation (1 mM NADH) by the SMP was 120 nmol O2 min?1mg?1, which was reduced to 98 nmol O2 min?1 (mg mitochondrial protein)?1 in the presence of EGTA. In contrast, the oxidation of NADH by intact mitochondria was completely inhibited by EGTA (from 182 to 14 nmol O2 min?1mg?1). The EGTA-resistant NADH oxidation by the SMP is ascribed to the NADH dehydrogenase(s) on the inside of the inner membrane and exposed to the medium in the inside-out SMP. In the presence of EGTA it could be shown that two NADH dehydrogenase activities were present in the SMP. One had an apparent Km of 7 μM for NADH, a Vmax of 80 nmol NADH min?1mg?1, and was rotenone-sensitive. This dehydrogenase is equivalent to the mammalian Complex I NADH dehydrogenase. The other dehydrogenase, which was rotenone-resistant, had a Km of 80 μM and a Vmax of 131 nmol NADH min?1mg?1; it is probably responsible for the rotenone-resistant oxidation of organic acids often observed in plant mitochondria. The redox poise of the pyridine nucleotides had only a small effect on the relative rates of the two internal dehydrogenases. Electron flow through these dehydrogenases appears, therefore, to be regulated mainly by the concentration of NADH in the matrix of the mitochondria.  相似文献   

20.
A dual-tube dense-phase silicone rubber membrane bioreactor was investigated for control of cyclohexane-contaminated air as part of a jet propulsion (JP-8) fuel remediation investigation strategy. The reactor was seeded with a mixed bacterial consortium isolated from the water/fuel interface of a JP-8 jet fuel sample and activated sludge, capable of aromatic and cyclic compound biodegradation. Cyclohexane removal ranged from 1.1 to 28.6 mg L?1, with removal percentages ranging from 4.6% to 37.6%. Removal in the bioreactor ranged from 29.4 to 596.6 mg min?1 m?2 and measured elimination capacities ranged from 46.7 to 947.9 g m?3 h?1. Removal rates and elimination capacity increased with increasing biofilm growth and with increasing loading rates of cyclohexane. Loading rates ranged from 395.9 to 2189.5 mg min?1 m?2. Results of this study showed effective removal of cyclohexane using the membrane bioreactor, suggesting that this technology may have applicability for treating vapors contaminated with cyclic hydrocarbons.  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