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 functional significance of the apical vacuolar-type proton pump (V-ATPase) in Drosophila Malpighian tubules was studied by measuring the intracellular pH (pHi) and luminal pH (pHlu) with double-barrelled pH-microelectrodes in proximal segments of the larval anterior tubule immersed in nominally bicarbonate-free solutions (pHo 6.9). In proximal segments both pHi (7.43±0.20) and pHlu (7.10±0.24) were significantly lower than in distal segments (pHi 7.70±0.29, pHlu 8.09±0.15). Steady-state pHi of proximal segments was much less sensitive to changes in pHo than pH of the luminal fluid (pHlu/pHo was 0.49 while pHi/pHo was 0.18; pHo 6.50–7.20). Re-alkaliniziation from an NH4Cl-induced intracellular acid load (initial pHi recovery rate 0.55±0.34 pH·min-1) was nearly totally inhibited by 1 mmol·l-1 KCN (96% inhibition) and to a large degree (79%) by 1 mol·l-1 bafilomycin A1. In contrast, both vanadate (1 mmol·l-1) and amiloride (1 mmol·l-1) inhibited pHi recovery by 38% and 33%, respectively. Unlike amiloride, removal of Na+ from the bathing saline had no effect on pHi recovery, indicating that a Na+/H+ exchange is not significantly involved in pHi regulation. Instead pHi regulation apparently depended largely on the availability of ATP and on the activity of the bafilomycin-sensitive proton pump.Abbreviations DMSO dimethylsulphoxide - DNP 2,4-dinitrophenol - NMDG N-methyl-D-glucamine - pHi intracellular pH - pHlu pH of the luminal fluid - pHo pH of the superfusion medium - I intrinsic intracellular buffer capacity  相似文献   

2.
Summary The intrinsic viscosity of phosphofructokinase fromDunaliella salina in different states of aggregation was determined. The instrinsic viscosity [], of the biologically active tetramer, with a molecular weight of 320,000, was found to be 6.5 ml·g–1 at 4°C. Moreover, for the inactive dimer, with a molecular weight of 160,000, a value of []=8.0 ml·g–1 was determined. The high molecular weight aggregate of phosphofructokinase fromDunaliella salina, that shows little activity, has an intrinsic viscosity of 23.2 ml·g–1, which is significantly higher than that found for the active tetramer and the inactive dimer.Small angle X-ray scattering experiments in solution of this high molecular from of phosphofructokinase fromDunaliella salina reveal a radius of gyration of the cross section ofR c=49.0 Å at an ionic strength of 0.15 M andpH 7.2. Furthermore, a comparison of the values obtained for the tetramer and the radius of gyration (R g=52.9 Å) with those of typical spherical proteins (3–4 ml·g–1) shows that the values of [] andR g are significantly larger for the high molecular weight form of phosphofructokinase than for the spherical proteins. The high intrinsic viscosity of the polymeric form of phosphofructokinase suggests an end-to-end aggregation consisting of monomeric units with heights,h=80–90 Å, and a cylindrical diameter of approximately 140.0 Å, resulting in a long rod of a total length of 1,800 Å and a molecular weight of two million. On the basis of the experimentally observedR c and [] values, using a prolate ellipsoid of revolution as a model, the hydrodynamic volume and the hydration, the axial ratio could be determined to be 12. The native tetrameric form contains 0.4 g H2O/g protein, whereas the higher aggregate structure corresponds to a hydration of 0.60 g H2O/g protein.  相似文献   

3.
Parallel experiments on human tooth enamel with sp.g. >2.95 and synthetic hydroxyapatite have been carried out to study the substitution of CO 3 2– for OH, produced at 1000°C in dry CO2, with the complementary use of neutron diffraction, x-ray diffraction, infrared spectroscopy and thermochemical techniques. It was verified that the substitution (i) is CO 3 2– 2(OH and is completely reversible on specimen exposure to H2O vapor at 1000°C, (ii) takes place with the carbon atom on or near the hexad axis, (ii) places one CO 3 2– group per unit cell in an ordered fashion and so changes the space group fromP6 3/m to one without a screw axis, (iii) was consistent, by its incompleteness, with the occurrence of substitution of O2– for 2(OH) in 25 to 40% of the unit cells, (iv) produced similar marked changes in the neutron powder diffraction patterns of both tooth enamel and hydroxyapatite, and (v) took place at a site where only a minor portion of the CO 3 2– in normal untreated human tooth enamel occurs. From comparative Rietveld analysis results from x-ray and neutron powder diffraction patterns it is suggested that the carbon atom of this A-site CO 3 2– is near 0,0,0.12 and the CO 3 2– plane makes an angle of 18° with thec direction.On being heated at 400°C in H2O vapor, tooth enamel retained much of its CO 3 2– but its a lattice parameter changed from 9.445(3) Å to 9.420(1) Å, essentially that of hydroxyapatite. After once being heated at high temperatures, tooth enamel and hydroxyaptite showed similar responses to various treatments, including carbonation. After heating, more -Ca3(PO4)2 was found in the tooth enamel specimen. Comparative weight change, IR, and other data for tooth enamel and hydroxyapatite heated in He, then in CO2, and then in H2O vapor showed a 20% or more deficiency of structural OH in the untreated tooth enamel.  相似文献   

4.
Summary Active transport of potassium in K+-starvedNeurospora was previously shown to resemble closely potassium uptake in yeast,Chlorella, and higher plants, for which K+ pumps or K+/H+-ATPases had been proposed. ForNeurospora, however, potassium-proton cotransport was demonstrated to operate, with a coupling ratio of 1 H+ to 1 K+ taken inward so that K+, but not H+, moves against its electrochemical gradient (Rodriguez-Navarro et al.,J. Gen. Physiol. 87:649–674).In the present experiments, the current-voltage (I–V) characteristic of K+–H+ cotransport in spherical cells ofNeurospora has been studied with a voltage-clamp technique, using difference-current methods to dissect it from other ion-transport processes in theNeurospora plasma membrane. Addition of 5-200 M K+ to the bathing medium causes 10–150 mV depolarization of the unclamped membrane, and yields a sigmoidI–V curve with a steep slope (maximal conductance of 10–30 S/cm2) for voltages of –300 to –100 mV, i.e., in the normal physiologic range. Outside that range the apparentI–V curve of the K+-H+ symport saturates for both hyperpolarization and depolarization. It fails to cross the voltage axis at its predicted reversal potential, however, an effect which can be attributed to failure of theI–V difference method under reversing conditions.In the absence of voltage clamping, inhibitors—such as cyanide or vanadate—which block the primary proton pump inNeurospora also promptly inhibit K+ transport and K+-H+ currents. But when voltage clamping is used to offset the depolarizing effects of pump blockade, the inhibitors have no immediate effect on K+-H+ currents. Thus, the inhibition of K+ transport usually observed with these agents reflects the kinetic effect of membrane depolarization rather than any direct chemical action on the cotransport system itself.Detailed study of the effects of [K+]o and pHo on theI–V curve for K+-H+ symport has revealed that increasing membrane potential systematicallydecreases the apparent affinity of the transporter for K+, butincreases affinity for protons (K m range: for [K+]o, 15–45 M; for [H+]o, 10–35 nM). This behavior is consistent with two distinct reaction-kinetic models, in which (i) a neutral carrier binds K+ first and H+ last in the forward direction of transport, or (ii) a negatively charged carrier (–2) binds H+ first and K+ last.  相似文献   

5.
New 3D HCN quantitative J (QJ) pulse schemes are presented for the precise and accurate measurement of one-bond 15N1/913C1, 15N1/913C6/8, and 15N1/913C2/4 residual dipolar couplings (RDCs) in weakly aligned nucleic acids. The methods employ 1H–13C multiple quantum (MQ) coherence or TROSY-type pulse sequences for optimal resolution and sensitivity. RDCs are obtained from the intensity ratio of H1–C1–N1/9 (MQ-HCN-QJ) or H6/8–C6/8–N1/9 (TROSY-HCN-QJ) correlations in two interleaved 3D NMR spectra, with dephasing intervals of zero (reference spectrum) and 1/(2JNC) (attenuated spectrum). The different types of 15N–13C couplings can be obtained by using either the 3D MQ-HCN-QJ or TROSY-HCN-QJ pulse scheme, with the appropriate setting of the duration of the constant-time 15N evolution period and the offset of two frequency-selective 13C pulses. The methods are demonstrated for a uniformly 13C, 15N-enriched 24-nucleotide stem-loop RNA sequence, helix-35, aligned in the magnetic field using phage Pf1. For measurements of RDCs systematic errors are found to be negligible, and experiments performed on a 1.5 mM helix-35 sample result in an estimated precision of ca. 0.07 Hz for 1DNC, indicating the utility of the measured RDCs in structure validation and refinement. Indeed, for a complete set of 15N1/913C1, 15N1/913C6/8, and 15N1/913C2/4 dipolar couplings obtained for the stem nucleotides, the measured RDCs are in excellent agreement with those predicted for an NMR structure of helix-35, refined using independently measured observables, including 13C–1H, 13C–13C and 1H–1H dipolar couplings.Supplementary material to this paper is available in electronic form at http://dx.doi.org/10.1007/s10858-005-0646-2.  相似文献   

6.
Jones  M. B.  Humphries  S. W. 《Hydrobiologia》2002,488(1-3):107-113
Fluxes of CO2 and H2O vapour were measured by eddy covariance from a stand of the C4 emergent sedge Cyperus papyrus (papyrus), which formed a fringing swamp on the north-west shore of Lake Naivasha, Kenya. The fluxes of CO2 and H2O vapour between the papyrus swamp and the atmosphere were large but variable, depending on the hydrology of the wetland system and the condition of the vegetation. These measurements, combined with simulation modelling of annual fluxes of CO2, show that papyrus swamps have the potential to sequester large amounts of the carbon (1.6 kg C m–2 y–1) when detritus accumulates under water in anaerobic conditions, but they are a net source of carbon release to the atmosphere (1.0 kg C m–2 y–1) when water levels fall to expose detritus and rhizomes to aerobic conditions. Evapotranspiration from papyrus swamps (E) was frequently lower than evaporation from open water surfaces (E o) and plant factors have a strong influence on the flux of water to the atmosphere. For the period of measurement E/Eo was 0.36.  相似文献   

7.
In situ paired light and dark-stirred benthic flux chambers were used to estimate dissolved oxygen flux across the sediment–water interface in Lake Mývatn, Iceland. Three sampling stations were selected, each station reflecting a specific sedimentary environment, benthic communities, and water depth. During this study the phytoplankton density was low. Spatial and seasonal variations of bottom DO concentration and DO flux have been observed during this study. The oxygen consumption rate at all study sites had a mean of –89 (±44) mmol m–2 d–1 while the oxygen production rate due to benthic algae had a mean of 131 (±103) mmol m–2 d–1. There was a strong correlation (r=0.91) between oxygen consumption rate and temperature. This was presumably because of the temperature influence on rates of microbial and macrobenthic processes. The mean benthic primary production rate at all study sites was 1216 (±957) mg C m–2 d–1 between June 2000 and February 2001. Annual gross benthic primary production was estimated from the gross mean daily benthic DO production (P) and Redfield's C:O2 ratio of 106:138 to be 420 g C m–2 y–1 at station HO, 250 g C m–2 y–1 at B2 and 340 g C m–2 y–1 at station 95. Thus, the mean gross benthic primary production was estimated as 1151 mg C m–2 d–1 at station HO, 685 mg C m–2 d–1 at station B2, and 932 mg C m–2 d–1 at station 95.  相似文献   

8.
Summary Mechanisms of proton conductance (G H) were investigated in phospholipid bilayer membranes containing long-chain fatty acids (lauric, myristic, palmitic, oleic or phytanic). Membranes were formed from diphytanoyl phosphatidylcholine in decane plus chlorodecane (usually 30% vol/vol). Fatty acids were added either to the aqueous phase or to the membrane-forming solution. Proton conductance was calculated from the steadystate total conductance and the H+ diffusion potential produced by a transmembrane pH gradient. Fatty acids causedG H to increase in proportion to the first power of the fatty acid concentration. TheG H induced by fatty acids was inhibited by phloretin, low pH and serum albumin.G H was increased by chlorodecane, and the voltage dependence ofG H was superlinear. The results suggest that fatty acids act as simple (A type) proton carriers. The membrane: water partition coefficient (K p ) and adsorption coefficient () were estimated by finding the membrane and aqueous fatty acid concentrations which gave identical values ofG H. For palmitic and oleic acidsK p was about 105 and was about 10–2 cm. The A translocation or flip-flop rate (k a ) was estimated from the value ofG H and the fatty acid concentration in the membrane, assuming that A translocation was the rate limiting step in H+ transport. Thek A 's were about 10–4 sec–1, slower than classical weak-acid uncouplers by a factor of 105. Although long-chain fatty acids are relatively inefficient H+ carriers, they may cause significant biological H+ conductance when present in the membrane at high concentrations, e.g., in ischemia, hypoxia, hormonally induced lipolysis, or certain hereditary disorders, e.g., Refsum's (phytanic acid storage) disease.  相似文献   

9.
Summary The course of the CO2 evolution rates of soil samples has been followed continuously in the absence and in the presence of various organic compounds. After an incubation period of 300 hours at 13 and 20°C the CO2 evolution from pasture soil (containing 1.76% soil organic carbon) amounted to 0.13 and 0.44g CO2–C.g soil–1.h–1, respectively. For arable soil (containing 1.20% soil organic carbon) the rates amounted to 0.04 and 0.09 g CO2–C.g soil–1.h–1, respectively.At 20°C larger amounts of the organic substrates added to the soil supplied with 20 g NH4NO3–N.g soil–1 were lost as CO2 than at 13°C, indicating a higher efficiency of the growth of microorganisms at lower temperatures. In the absence of NH4NO3 the respiration rates were initially higher than in its presence, suggesting that a part of the soil microflora is inhibited by low concentrations of NH4NO3. The amounts of carbon lost were low for phenolcarboxylic acids with OH groups in the ortho position. The replacement of one of these groups by a methoxyl group resulted in a larger amount of the C lost as CO2. The replacement of the COOH group by a C=C–COOH group had a decreasing effect on the decomposition of the phenolic acids tested. The decomposition of vanillic acid,p-hydroxybenzoic acid, and of the benzoic acids with OH groups in the meta position was as complete as that of glucose, amino acids or casein. The decomposition of bacterial cells to CO2 was considerably less than that of glucose.No evidence could be obtained that the low percentage of substrate converted to CO2 at the time of maximal respiration rate was due to the decreasing diffusion rate of substrate to the microbial colonies in the soil during the consumption of substrate.  相似文献   

10.
Summary Possible reactions of thiyl free radicals in biological environment are reviewed. In particular hydrogen transfer processes from model C–H compounds like alcohols and ethers as well as from polyunsaturated fatty acids to thiyl radicals are described to proceed with reasonably high rate constants (103 – 104 and 106 – 107 M–1 s–1, respectively). Thiyl radicals have thus to be considered as potentially hazardous species especially with respect to DNA damage and lipid peroxidation.Paper given at the workshop Molecular Radiation Biology. German Section of the DNA Repair Network, München-Neuherberg, 21.–23.3.1990  相似文献   

11.
Penetration of 3H-labeled water (3H2O) and the 14C-labeled organic acids benzoic acid ([14C]BA), salicylic acid ([14C]SA), and 2,4-dichlorophenoxyacetic acid ([14C]2,4-D) were measured simultaneously in isolated cuticular membranes of Prunus laurocerasus L., Ginkgo biloba L., and Juglans regia L. For each of the three pairs of compounds (3H2O/[14C]BA, 3H2O/[14C]SA, and 3H2O/[14C]2,4-D) rates of cuticular water penetration were highly correlated with the rates of penetration of the organic acids. Therefore, water and organic acids penetrated the cuticles by the same routes. With the combination 3H2O/[14C]BA, co-permeability was measured with isolated cuticles of nine other plant species. Permeances of 3H2O of all 12 investigated species were highly correlated with the permeances of [14C]BA (r2 = 0.95). Thus, cuticular transpiration can be predicted from BA permeance. The application of this experimental method, together with the established prediction equation, offers the opportunity to answer several important questions about cuticular transport physiology in future investigations.  相似文献   

12.
Summary For numerical solution of the reaction-mass transfer equations for immobilised biocatalysts it may be better to start integration at the particle surface and proceed inwards: calculations are targetted on the region to which practically interesting changes are often confined (because concentrations are effectively zero in the interior); and during iterative solution wrong initial estimates may be rejected after detecting anomalies early in the integration.Symbols Cb substrate concentration in bulk (mol m–3) - c dimensionless substrate concentration (C/Cb) (-) - De effective diffusion coefficient (m2s–1) - Da Damkohler number (V.ro 2/De.Ks) (-) - Ks substrate concentration kinetic coefficient (mol m–3) - ke external mass transfer coefficient (ms–1) - ro bead radius (m) - Sh Sherwood number (ke.ro/De) (-) - V maximum rate per unit volume in beads (mol m–3s–1) - x dimensionless distance from bead centre (r/ro) (-) - dimensionless kinetic coefficient (Ks/Cb) (-) - o effectiveness factor (-)  相似文献   

13.
We have applied enzyme kinetic analysis to electrophysiological steady-state data of Zhou et al. (Zhou, J.J., Trueman, L.J., Boorer, K.J., Theodoulou, F.L., Forde, B.G., Miller, A.J. 2000. A high-affinity fungal nitrate carrier with two transport mechanisms. J. Biol. Chem. 275:39894–9) and to new current-voltage-time records from Xenopus oocytes with functionally expressed NrtA (crnA) 2H+-NO 3 symporter from Emericella (Aspergillus) nidulans. Zhou et al. stressed two Michaelis-Menten (MM) mechanisms to mediate the observed nitrate-induced currents, I NO 3 . We show that a single straightforward reaction cycle describes the data well, pointing out that during exposure to external substrate, S = (2H++NO 3 )o, the product concentration inside, [P] = [H+] i 2 · [NO 3 i, may rise substantially near the plasma membrane, violating the condition [P] [S] for MM kinetics. Here, [P] and its changes during experimentation are treated explicitly. K 1/2 20 µM for I NO 3 at pHo from Zhou et al. is confirmed. According to our analysis, NrtA operates between about 0.2 and 0.6 of the electrical distance in the membrane (outside 0, inside 1). In absence of thermodynamic gradients, the predominant orientation of the binding site(s) is probably inwards. The activity of the enzyme is sensitive to the transmembrane voltage, V, with an apparent gating charge of +1.0 ± 0.5 for inactivation, and transition probabilities of 0.3–1.3 s–1 at V = 0. This gating mode impedes loss of cellular NO 3 during depolarization.  相似文献   

14.
Compartmentation and flux characteristics of nitrate in spruce   总被引:8,自引:0,他引:8  
The radiotracer13N was used to undertake compartmental analyses for NO 3 in intact non-mycorrhizal roots ofPicea glauca (Moench) Voss. seedlings. Three compartments were defined, with half-lives of exchange of 2.5 s, 20 s, and 7 min. These were identified as representing surface adsorption, apparent free space, and cytoplasm, respectively. Influx, efflux, and net flux as well as cytoplasmic and apparent-free-space nitrate concentrations were estimated for three different concentration regimes of external nitrate. After exposure to external NO 3 for 3 d, influx was calculated to be 0.09 mol·g–1·h–1 (at 10 M [NO 3 ]o), 0.5mol·g–1·h–1 (at 100 M [NO inf3 sup– ]o), and 1.2 mol · g–1· h–1 (at 1.5 mM [NO 3 ]o). Efflux increased with increasing [NO 3 ]o, constituting 4% of influx at 10 M, 6% at 100 M, and 21% at 1.5 mM. Cytoplasmic [NO 3 ] was estimated to be 0.3 mM at 10 uM [NO 3 ]o, 2mM at 100 M [NO 3 ]o, and 4mM at 1.5 mM [NO 3 ]o, while free-space [NO 3 ] was 16 M, 173 M, and 2.2 mM, respectively. A series of experiments was carried out to confirm the identity of the compartments resolved by efflux analysis. Pretreatment at high temperature or application of 2-chloro-ethanol, sodium dodecyl sulphate or hydrogen peroxide made it possible to distinguish the metabolic (cytoplasmic) phase from the remaining two (physical) phases. Likewise, varying [Pi] of the medium altered efflux and thereby [NO 3 ]cyt, but did not affect [NO 3 ]free space.Abbreviations and Symbols [NO 3 ]cyt cytoplasmic NO 3 concentration - [NO 3 ]free space apparent-free-space NO 3 concentration - [NO 3 ]o concentration of NO 3 in the external solution - NO 3 flux - co efflux from the cytoplasm - oc influx to the cytoplasm - net net flux - xylem flux to the xylem - red/vac combined flux to reduction and the vacuole The research was supported by a Natural Sciences and Engineering Research Council, Canada, grant to Dr. A.D.M. Glass and by a University of British Columbia Graduate Fellowship to Herbert J. Kronzucker. Our thanks go to Dr. M. Adam and Mr. P. Culbert at the particle accelerator facility TRIUMF on the University of British Columbia Campus for providing13NO 3 , Drs. R.D. Guy and S. Silim for providing plant material, and Dr. M.Y. Wang, Mr. J. Mehroke and Mr. P. Poon for assistance in experiments and for helpful discussions.  相似文献   

15.
Production of hydrogen peroxide has been found in Ulva rigida (Chlorophyta). The formation of H2O2 was light dependent with a production of 1.2 mol·g FW–1·h–1 in sea water (pH 8.2) at an irradiance of 700 mol photons m–2·s–1. The excretion was also pH dependent: in pH 6.5 the production was not detectable (< 5 nmol·g FW–1·h–1) but at pH 9.0 the production was 5.0 mol·g FW–1·h–1. The production of H2O2 was totally inhibited by 3-(3,4-dichlorophenyl)-1,1 dimethylurea (DCMU). The ability of U. rigida growing in tanks (7501) under a natural light regime to excrete H2O2 was checked and found to be seven times higher at 08.00 hours than other times of the day. The H2O2 concentration in the cultivation tank (density: 2 g FW·l–1) reached the highest value (3 M) at 11.00 hours. Photosynthesis was not influenced by H2O2 formation. The H2O2 is suggested to come from the Mehler reaction (pseudocyclic photophosphorylation). With an oxygen evolution of 120 mmol·g FW–1·h–1 at pH 8.2 and 90 mmol·g FW–1·h–1 at pH 9.0, 0.5% and 2.7% of the electrons were used for extracellular H2O2 production. The H2O2 production is sufficiently high to be of physiological and ecological significance, and is suggested to be a part of the defence against epi and endophytes.Abbreviations ACL artificial, continuous light - DCMU 3-(3,4-dichlorophenyl)-1,1-dimethylurea - GNL greenhouse - LDC Luminol-dependent chemiluminescence - SOD Superoxide dismutase This investigation was supported by SAREC (Swedish Agency for Research Cooperation with Developing Countries), Hierta-Retzius Foundation, Marianne and Marcus Wallenberg Foundation, the Swedish Environmental Protection Board, and CICYT Spain.  相似文献   

16.
Summary The apical membrane of the rabbit corneal endothelium contains a potassium-selective ionic channel. In patch-clamp recordings, the probability of finding the channel in the open state (P o) depends on the presence of either HCO 3 or Cl in the bathing medium. In a methane sulfonate-containing bath,P o is <0.05 at all physiologically relevant transmembrane voltages. With 0mm [HCO 3 ] o at +60 mV,P o was 0.085 and increased to 0.40 when [HCO 3 ] o was 15mm. With 4mm [Cl] o at +60 mV,P o was 0.083 and with 150mm Cl,P o increased to 0.36. LowP o's are also found when propionate, sulphate, bromide, and nitrate are the primary bath anions. The mechanism of action of the anion-stimulated K+ channel gating is not yet known, but a direct action of pH seems unlikely. The alkalinization of cytoplasm associated with the addition of 10mm (NH4)2SO4 to the bath and the acidification accompanying its removal do not result in channel activation nor does the use of Nigericin to equilibrate intracellular pH with that of the bath over the pH range of 6.8 to 7.8. Channel gating also is not affected by bathing the internal surface of the patch with cAMP, cGMP, GTP--s, Mg2+ or ATP. Blockers of Na/H+ exchange, Na+–HCO 3 cotransport, Na+–K+ ATPase and carbonic anhydrase do not block the HCO 3 stimulation ofP o. Several of the properties of the channel could explain some of the previously reported voltage changes that occur in corneal endothelial cells stimulated by extracellular anions.  相似文献   

17.
Packed-bed bioreactors containing activated carbon as support carrier were used to produce H2 anaerobically from a sucrose-limiting medium while acclimated sewage sludge was used as the H2 producer. The effects of bed porosity (b) and substrate loading rate on H2 fermentation were examined using packed beds with b of 70–90% being operated at hydraulic retention times (HRT) of 0.5–4 h. Higher b and lower HRT favored H2 production. With 20 g COD l–1 of sucrose in the feed, the optimal H2 production rate (7.4 l h–1 l–1) was obtained when the bed with b=90% was operated at HRT = 0.5 h. Flocculation of cells enhanced the retention of sludge for stable operations of the bioreactor at low HRTs. The gas products resulting from fermentative H2 production consisted of 30–40% H2 and 60–70% CO2. Butyric acid was the primary soluble product, followed by propionic acid and valeric acid.  相似文献   

18.
Summary Horseradish peroxidase C (HRP; ferric) reacts with H2O2 to form Compound I, with an equilibrium constant of about 1014 M–1. Two-step reduction of Compound I to Compound II and further to the ferric enzyme occurs reversibly at Eo values of 0.90 and 0.93 V (pH 7.0), respectively. The pH dependence of Eo values for each one-electron step, ferrous ferric Compound II Compound I indicates the presence of redox-linked ionization at pKa values of 7.3 in the ferrous state, 11.0 in the ferric and 8.6 in Compound II. Zinc-substituted HRP C is oxidized to its free-radical form at an Eo value of 0.74 (pH 6.0). Comparison of oxidized zinc HRP C with Compound I shows that Compound I contains a porphyrin -cation radical. The flash photolysis study on the NO-ferric HRP C complex clearly indicates that the iron is pentacoordinated in HRP C while it is hexacoordinated in metmyoglobin. From the kinetic analysis of the acid-alkaline conversion of HRP C, the second-order rate constants of the reactions with H+ and HO are estimated to be 1.5 × 1010 and 6.7 × 104 M–1s–1, respectively. The latter rate constant greatly varies with the kind of hemoproteins. In the presence of HRP C and O2, indole-3-acetate is oxidized to its hydroperoxide form, which reacts effectively with HRP C to form Compound I and further converts Compound I to a verdohemoprotein.Abbreviations HRP horseradish peroxidase (HRP without subgroup letter denotes a classical preparation consisting of HRP B and HRP C) - EPR electron paramagnetic resonance - NMR nuclear magnetic resonance  相似文献   

19.
Summary The linear growth phase in cultures limited by intracellular (conservative) substrate is represented by a flat exponential curve. Within the range of experimental errors, the presented model fits well the data from both batch and continuous cultures ofEscherichia coli, whose growth is limited in that way.List of symbols D dilution rate, h–1 - KS saturation constant, g.L–1 - S concentration of the limiting substrate, g.L–1 - Si concentration of the limiting substrate accumulated in the cells, g.g–1 - So initial concentration of the limiting substrate, g.L–1 - t time of cultivation, h - t1 time of exhaustion of the limiting substrate from medium, h - to beginning of exponential phase, h - X biomass concentration, g.L–1 - X1 biomass concentration at the time of exhaustion of the limiting substrate from the medium, g.L–1 - Xo biomass concn. at the beginning of exponential phase, g.L–1 - biomass concn. at steady-state, g.L–1 - Y growth yield coefficient (biomass/substrate) - specific growth rate, h–1 - m maximum specific growth rate, h–1  相似文献   

20.
Granum  Espen  Myklestad  Sverre M. 《Hydrobiologia》2002,477(1-3):155-161
A new method is described for the combined determination of -1,3-glucan and cell wall polysaccharides in diatoms, representing total cellular carbohydrate. The glucan is extracted by 0.05 mol l–1 H2SO4 at 60 °C for 10 min, and the cell wall polysaccharides are subsequently hydrolyzed by 80% H2SO4 at 0–4 °C for 20 h. Each carbohydrate fraction is determined by the phenol-sulphuric acid method. The method has been demonstrated for axenic cultures of the marine diatom Skeletonema costatum and natural marine phytoplankton populations dominated by diatoms. Cellular glucan and cell wall polysaccharides were determined with standard deviations of 1–3% and 2–5%, respectively.  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