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 468 毫秒
1.
Summary Resting cells ofArthrobacter sp. (DSM 3745) with the ability to form L-tryptophan from D,L-5-(3-indolylmethy)hydantoin were used for the bioconversion of D,L-5-- and D,L-5--naphthylmethylhydantoin (D,L-5-- and D,L-5--NMH) to the corresponding L-amino acids. Under the optimal reaction conditions of pH 9.7 and 40°C specific productivities of 0.2 (-naphtylalanine) and 0.6 (-naphtylalanine) mM amino acid x g cell dry mass–1 x h–1 were obtained in a 0.1 M Na2CO3/NaHCO3-buffer in a strirred bioreactor.  相似文献   

2.
AxenicTrentepohlia odorata was cultured at three different NH4Cl levels (3.5 × 10–2, 3.5 × 10–3, 3.5 × 10–4 M) and three different light intensities (48, 76, 122 µmol m–2 s–1). Chloride had no effect on growth over this range of concentration. High light intensity and high NH4Cl concentration enhanced the specific growth rate. The carotenoid content increased under a combination of high light intensity and low N concentration. WhenD. bardawil was exposed to the same combination of growth conditions, there was an increase in its carotenoid content. The light saturation and the light inhibition constants (K s andK i, respectively) for growth, and the saturation constant (K m) for NH4Cl were determined. TheK s andK i values were higher inT. odorata (66.7 and> 122 mol m–2 s–1, respectively) than inD. bardawil (5.1 and 14.7 µmol m–2 s–1, respectively). TheK m value determined at 122 µmol m–2 s–1, however, was lower inT. odorata (0.048 µM) than inD. bardawil (0.062 µM).Author for correspondence  相似文献   

3.
Eicosapentaenoic (EPA) and docosahexaenoic (DHA) acid productivities from chemostat cultures of an isolate of Isochrysis galbana have been studied. The productivities reached in the interval of dilution rates between 0.0295 h–1 and 0.0355 h–1 were 1.5mg·1–1·h–1 for lipids, 300 g·1–1·h–1 for EPA and 130g1·1–1·h–1 for DHA. Furthermore, light attenuation by mutual shading, and agitation speed influences on growth and fatty acid composition were analysed. A model relating steady-state dilution rates to internal average light intensity has been proposed, the parameter values of which obtained by non-linear regression were: maximum specific growth rate (max)=0.0426 h–1; the affinity of cells to light (Ik) = 10.92 W·m–2; the exponent (n) = 5.13; regression coefficient (r 2)=0.9999. Correspondence to: E. Molina Grima  相似文献   

4.
Biochemical and biophysical parameters, including D1-protein turnover, chlorophyll fluorescence, oxygen evolution activity and zeaxanthin formation were measured in the marine seagrassZostera capricorni (Aschers) in response to limiting (100 mol·m–2·–1), saturating (350 mol·m–2·s–1) or photoinhibitory (1100 mol·m–2·s–1) irradiances. Synthesis of D1 was maximal at 350 mol·m–2·s–1 which was also the irradiance at which the rate of photosynthetic O2 evolution was maximal. Degradation of D1 was saturated at 350 mol·m–2·s–1. The rate of D1 synthesis at 1100 mol·m–2·s–1 was very similar to that at 350 mol·m–2·s–1 for the first 90 min but then declined. At limiting or saturating irradiance little change was observed in the ratio of variable to maximal fluorescence (Fv/Fm) measured after dark adaptation of the leaves, while significant photoinhibition occurred at 1100 mol·m–2·s–1. The proportion of zeaxanthin in the total xanthophyll pool increased with increasing irradiance, indicative of the presence of a photoprotective xanthophyll cycle in this seagrass. These results are consistent with a high level of regulatory D1 turnover inZostera under non-photoinhibitory irradiance conditions, as has been found previously for terrestrial plants.We would like to thank Professor Peter Böger (Department of Plant Biochemistry, University of Konstanz, Germany) for the kind gift of D1 antibodies. This work was partly supported by a University of Queensland Enabling Grant to CC.  相似文献   

5.
Densities ofAmblyomma americanum (L.) onBos indicus, B. taurus andB. indicus x B. taurus cattle are compared over a 3-year period, and the growth rate (rate of increase or decrease) of parasitic tick populations on each cattle genotype is estimated.Average log10 densities of parasiticA. americanum larvae are significantly (P=0.05) lower onB. indicus cattle than onB. taurus andB. indicus x B. taurus cattle. Average log densities of nymphal and adult ticks onB. taurus cattle are significantly higher than onB. indicus cattle but neither cattle genotype differs in this regard fromB. indicus x B. taurus cattle.Estimated annual tick population growth rates (log10) for parasiticA. americanum are positive onB. taurus cattle (+0.84 larvae, +0.09 nymphs, +0.22 adults calf–1 year–1), but are negative onB. indicus (–0.18 nymphs, –0.14 adults calf–1 year–1) andB. indicus x B. taurus cattle (–0.45 larvae, –0.24 nymphs, –0.14 adults calf–1 year–1). Populations of parasitic larvae were not detected onB. indicus cattle.  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.
Hansen  Jonas  Reitzel  Kasper  Jensen  Henning S.  Andersen  Frede Ø. 《Hydrobiologia》2003,492(1-3):139-149
The effects of oxygen, aluminum, iron and nitrate additions on phosphate release from the sediment were evaluated in the softwater Lake Vedsted, Denmark, by a 34-day experiment with undisturbed sediment cores. Six treatments were applied: (1) Control - O2 (0–20% saturation), (2) O2 (100% saturation) (3) Al3+ – O2, (4) Fe3+ + O2, (5) Fe3+ – O2, and (6) NO3 – O2. Al2(SO4)3*18 H2O and FeCl3*4H2O were added in amounts that theoretically should immobilize the exchangeable P-pool in the top 5 cm of the sediment, while sodium nitrate concentrations were increased to 5 mg N l–1. The four treatments with metals or NO3 reduced the P efflux from the sediment significantly as compared to the suboxic control treatment. Mean accumulated P-release rates for suboxic treatments with Al3+, Fe3+, and NO3 were: –0.27 mmol m–2 (st. dev = 0.02 mmol m–2, N = 5), 0.58 mmol m–2 (st. dev = 0.30 mmol m–2, N = 5) and 1.40 mmol m–2 (st. dev = 0.14 mmol m–2, N = 5), respectively. The oxic treatment with Fe3+ had a P efflux of 0.36 mmol m–2 (st. dev = 0.08 mmol m–2, N = 5). The two highest P-release rates were observed in the control treatment and the treatment with O2 (14.50 mmol m–2 (st. dev = 3.90 mmol m–2, N = 5) and 2.31 mmol m–2 (st. dev = 0.80 mmol m–2, N = 5), respectively). In order to identify changes in the P and Fe binding sites in the sediment as caused by the treatments, a sequential P extraction procedure was applied on the sediment before and after the efflux experiment. Addition of O2, Fe3+ and NO3 to the sediment increased the amounts of oxidized Fe3+ and PBD. Al3+ addition resulted in a lower fraction of PBD but a correspondingly higher fraction of Al-bound P. Addition of Al3+ decreased the Fe-efflux from the suboxic sediment as well as the amount of oxidized Fe3+ in the sediment. This questions the use of Al compounds that contain sulfate because of the possible formation of FeS, which will restrict upward migration of Fe2+ and the formation of new Fe-oxides in the surface sediment. Instead, we suggest the use of AlCl3 for lake restoration purposes.  相似文献   

8.
High molecular weight lectins (> 100 kDa) from seeds of the legumes Canavalia brasiliensis (CnBr), Cratylia floribunda (CFL), Phaseolus vulgaris (PHA) and Vatairea macrocarpa (VML), temporarily stimulate the respiration of Rhizobium tropici-CIAT899 and R. etli-CFN42. These stimulants were significant (P < 0.05) in bacterial suspensions (> 2.85 mg dry biomass ml–1), having at least 6200 molecules of lectins per bacteria. The VML (20 g ml–1), induced specific O2 demand of 2.3–2.5 M O2 min–1 mg dry biomass–1, in CFN42 and CIAT899, respectively. However, CnBr, CFL and PHA induced smaller demands of O2 (5×), in both strains. The order of affinities of the lectins was approximately VML > PHA > CFL > CnBr, with regard to respiratory stimuli in CIAT899 strain. The co-administration of 10 g VML ml–1 and 9.8 M galactose, in CIAT899 suspensions, reduced the respiratory stimuli significantly in relation to the treatment with VML alone. These respiratory stimuli, induced by the lectins, increase the significance of the interaction lectin × Rhizobium in terms of bacterial physiology. Its understanding could be important in relation to bacterial symbiotic behaviour.  相似文献   

9.
A fermentation medium based on millet (Pennisetum typhoides) flour hydrolysate and a four-phase feeding strategy for fed-batch production of baker's yeast,Saccharomyces cerevisiae, are presented. Millet flour was prepared by dry-milling and sieving of whole grain. A 25% (w/v) flour mash was liquefied with a thermostable 1,4--d-glucanohydrolase (EC 3.2.1.1) in the presence of 100 ppm Ca2+, at 80°C, pH 6.1–6.3, for 1 h. The liquefied mash was saccharified with 1,4--d-glucan glucohydrolase (EC 3.2.1.3) at 55°C, pH 5.5, for 2 h. An average of 75% of the flour was hydrolysed and about 82% of the hydrolysate was glucose. The feeding profile, which was based on a model with desired specific growth rate range of 0.18–0.23 h–1, biomass yield coefficient of 0.5 g g–1 and feed substrate concentration of 200 g L–1, was implemented manually using the millet flour hydrolysate in test experiments and glucose feed in control experiments. The fermentation off-gas was analyzed on-line by mass spectrometry for the calculation of carbon dioxide production rate, oxygen up-take rate and the respiratory quotient. Off-line determination of biomass, ethanol and glucose were done, respectively, by dry weight, gas chromatography and spectrophotometry. Cell mass concentrations of 49.9–51.9 g L–1 were achieved in all experiments within 27 h of which the last 15 h were in the fedbatch mode. The average biomass yields for the millet flour and glucose media were 0.48 and 0.49 g g–1, respectively. No significant differences were observed between the dough-leavening activities of the products of the test and the control media and a commercial preparation of instant active dry yeast. Millet flour hydrolysate was established to be a satisfactory low cost replacement for glucose in the production of baking quality yeast.Nomenclature C ox Dissolved oxygen concentration (mg L–1) - CPR Carbon dioxide production rate (mmol h–1) - C s0 Glucose concentration in the feed (g L–1) - C s Substrate concentration in the fermenter (g L–1) - C s.crit Critical substrate concentration (g L–1) - E Ethanol concentration (g L–1) - F s Substrate flow rate (g h–1) - i Sample number (–) - K e Constant in Equation 6 (g L–1) - K o Constant in Equation 7 (mg L–1) - K s Constant in Equation 5 (g L–1) - m Specific maintenance term (h–1) - OUR Oxygen up-take rate (mmol h–1) - q ox Specific oxygen up-take rate (h–1) - q ox.max Maximum specific oxygen up-take rate (h–1) - q p Specific product formation rate (h–1) - q s Specific substrate up-take rate (g g–1 h–1) - q s.max Maximum specific substrate up-take rate (g g–1 h–1) - RQ Respiratory quotient (–) - S Total substrate in the fermenter at timet (g) - S 0 Substrate mass fraction in the feed (g g–1) - t Fermentation time (h) - V Instantaneous volume of the broth in the fermenter (L) - V 0 Starting volume in the fermenter (L) - V si Volume of samplei (L) - x Biomass concentration in the fermenter (g L–1) - X 0 Total amount of initial biomass (g) - X t Total amount of biomass at timet (g) - Y p/s Product yield coefficient on substrate (–) - Y x/e Biomass yield coefficient on ethanol (–) - Y x/s Biomass yield coefficient on substrate (–) Greek letters Moles of carbon per mole of yeast (–) - Moles of hydrogen atom per mole of yeast (–) - Moles of oxygen atom per mole of yeast (–) - Moles of nitrogen atom per mole of yeast (–) - Specific growth rate (h–1) - crit Critical specific growth rate (h–1) - E Specific ethanol up-take rate (h–1) - max.E Maximum specific ethanol up-take rate (h–1)  相似文献   

10.
Individual plants from the BC1F6 and BC1F8 backcross progenies of barley-wheat [H. marinum subsp. gussoneanum Hudson (=H. geniculatum All.) (2n = 28) × T. aestivum L. (2n = 42)] and the BC1F6 progeny of their amphiploids were used to obtain alloplasmic euploid (2n = 42) lines L-28, L-29, and L-49 and alloplasmic telocentric addition (2n = 42 + 2t) lines L-37, L-38, and L-50. The lines were examined by genomic in situ hybridization (GISH), microsatellite analysis, chromosome C-banding, and PCR analysis of the mitochondrial 18S/5S repeat. Lines L-29 and L-49 were characterized by substitution of wild barley chromosome 7H1 for common wheat chromosome 7D. In line L-49, common wheat chromosomes 1B, 5D, and 7D were substituted with homeologous barley chromosomes. Lines L-37, L-38, and L-50 each contained a pair of telocentric chromosomes, which corresponded to barley chromosome arm 7H1L. All lines displayed heteroplasmy for the mitochondrial 18S/5S locus; i.e., both barley and wheat sequences were found. Original Russian Text ? N.V. Trubacheeva, E.D. Badaeva, I.G. Adonina, L.I. Belova, E.P. Devyatkina, L.A. Pershina, 2008, published in Genetika, 2008, Vol. 44, No. 1, pp. 81–89.  相似文献   

11.
Turnover and distribution of root exudates of Zea mays   总被引:1,自引:0,他引:1  
Decomposition and distribution of root exudates of Zea mays L. were studied by means of 14CO2 pulse labeling of shoots on a loamy Haplic Luvisol. Plants were grown in two-compartment pots, where the lower part was separated from the roots by monofilament gauze. Root hairs, but not roots, penetrated through the gauze into the lower part of the soil. The root-free soil in the lower compartment was either sterilized with cycloheximide and streptomycin or remained non-sterile. In order to investigate exudate distribution, 3 days after the 14C labeling, the lower soil part was frozen and sliced into 15, one-mm thick layers using a microtome. Cumulative 14CO2 efflux from the soil during the first 3 days after 14C pulse labeling did not change during plant growth and amounted to about 13–20% of the total recovered 14C (41–55% of the carbon translocated below ground). Nighttime rate of total CO2 efflux was 1.5 times lower than during daytime because of tight coupling of exudation with photosynthesis intensity. The average CO2 efflux from the soil with Zea mays was about 74 g C g–1 day–1 (22 g C m–2 day–1), although, the contribution of plant roots to the total CO2 efflux from the soil was about 78%, and only 22% was respired from the soil organic matter. Zea mays transferred about 4 g m–2 of carbon under ground during 26 days of growth. Three zones of exudate concentrations were identified from the distribution of the 14C-activity in rhizosphere profiles after two labeling periods: (1) 1–2 (3) mm (maximal concentration of exudates) 2) 3–5 mm (presence of exudates is caused by their diffusion from the zone 1); (3) 6–10 mm (very insignificant amounts of exudates diffused from the previous zones). At the distance further than 10 mm no exudates were found. The calculated coefficient of exudate diffusion in the soil was 1.9 × 10–7 cm2 s–1.  相似文献   

12.
The enzyme glucose oxidase (GO) was covalently immobilized onto a poly(vinyl alcohol) hydrogel, cross-linked with glutardialdehyde and a polyazonium salt. To compare the kinetic parameters of immobilized GO with the known kinetic parameters of soluble GO, the diffusion cell method was used.Between two compartments, containing solutions with different glucose concentrations, a GO-containing hydrogel membrane was placed. Simultaneous diffusion through and enzymatic reaction in the membrane occurred. In this way diffusional effects of the membrane could be eliminated from the effective kinetic parameters to yield the inherent kinetic parameters.It appeared that the enzymatic reaction is independent of the oxygen concentration at oxygen concentrations 0.22 mol m–3 (Michaelis constant for oxygen < 0.22 mol m–3). Further, the Michaelis constant for glucose does not change dramatically after immobilizing the enzyme. The maximal reaction rate is depending on the enzyme concentration. As the enzyme concentration in the membrane is not exactly known (mainly due to leakage of enzyme out of the membrane during membrane preparation), only an estimation of the turnover number can be made.The diffusion cell method is easy to carry out. Still, some recommendations can be made on the performance.List of Symbols g , 0x partition coefficient of glucose and oxygen, respectively - thickness of the wetted membrane (m) - A m surface area of membrane (m–2) - C constant (mol2 m–3) - c g , c 0x concentration of glucose and oxygen, respectively (mol m–3) - c g,0 c g, glucose concentration at the filter-paper/membrane interface next to compartment A and B, respectively (mol m–3) - c g, A c g, B glucose concentration in compartment A and B, respectively (mol m–3) - c GO glucose oxidase concentration (mol m–3) - D eff effective diffusion coefficient (m2 s–1) - D m , D sl diffusion coefficient in, respectively, the membrane and the solution layer (m2 s–1) - d dl , d df , d sl thickness of, respectively, the diffusion layer, the filter-paper and the solution layer (m) - h B initial slope of concentration versus time curve of compartment B (mol m–3 s–1) - J flux (mol m–2 s–1) - J 0 flux in the membrane at membrane/filter-paper interface next to compartment A and B, respectively (mol m–2 s–1) - J A , J B flux leaving compartment A and entering compartment B, respectively (mol m–2 s–1) - J m flux through the membrane (mol m–2 s–1) - k total mass transfer coefficient (m s–1) - k 1 , k 2 rate constant of a particular reaction step (m3 mol–1 s–1) - k–1, k–2 rate constant of a particular reaction step (s–1) - k cat (intrinsic) catalytic constant of turnover number (s–1) - k cat * inherent catalytic constant, determined by inserting D m (s–1) - k cat ** inherent catalytic constant, determined by inserting D eff (s–1) - k m (g) (intrinsic) Michaelis constant for glucose (mol m–3) - k m (o) (intrinsic) Michaelis constant for oxygen (mol m–3) - k m * (g) inherent Michaelis constant for glucose (mol m–3) - k m * (o) inherent Michaelis constant for oxygen (mol m–3) - m GO number of moles of GO present (mol) - P m permeability of glucose in the mebrane (m s–1) - P eff effective permeability (m s–1) - V volume (m3) - v 0 initial reaction velocity (mol m–3 s–1) - V max ** inherent maximal reaction velocity, determined by inserting Deff (mol m–3 s–1) - x distance (m)  相似文献   

13.
The concentration dependence of the influx ofl-lysine in excised roots ofArabidopsis thaliana seedlings was analyzed for the wild-type (WT) and two mutants,rlt11 andraec1, which had been selected as resistant to lysine plus threonine, and to S-2-aminoethyl-l-cysteine, respectively. In the WT three components were resolved: (i) a high-affinity, low-capacity component [K m = 2.2 M;V max = 23 nmol·(g FW)–1·h–1]; (ii) a low-affinity, high-capacity component [K m = 159 M;V max = 742 nmol·(g FW)–1·h–1]; (iii) a component which is proportional to the external concentration, with a constant of proportionalityk = 104 nmol·(g FW)–1 h–1];·mM–1. The influx ofl-lysine in the mutants was lower than in the WT, notably in the concentration range 0.1–0.4 mM, where it was only 7% of that in the WT. In both mutants the reduced influx could be fully attributed to the absence of the low-affinity (high-K m ) component. This component most likely represents the activity of a specific basic-amino-acid transporter, since it was inhibited by several other basic amino acids (arginine, ornithine, hydroxylysine, aminoethylcysteine) but not byl-valine. The high-affinity uptake ofl-lysine may be due to the activity of at least two general amino acid transporters, as it was inhibitable byl-valine, and could be further dissected into two components with a high affinity (K i = 1–5 M; and a low affinity (K i = 0.5–1mM) forl-valine, respectively. Therlt11 andraecl mutant have the same phenotype and the corresponding loci were mapped on chromosome 1, but it is not yet clear whether they are allelic.Abbreviations AEC S-2-aminoethyl-l-cysteine - K i equilibrium constant - WT wild-type  相似文献   

14.
An assay is described that allows the direct measurement of the enzyme activity catalyzing the transfer of the methyl group from N 5-methyltetrahydromethanopterin (CH3–H4MPT) to coenzyme M (H–S–CoM) in methanogenic archaebacteria. With this method the topology, the partial purification, and the catalytic properties of the methyltransferase in methanol- and acetate-grown Methanosarcina barkeri and in H2/CO2-grown Methanobacterium thermoautotrophicum were studied. The enzyme activity was found to be associated almost completely with the membrane fraction and to require detergents for solubilization. The transferase activity in methanol-grown M. barkeri was studied in detail. The membrane fraction exhibited a specific activity of CH3–S–CoM formation from CH3–H4MPT (apparent K m=50 M) and H–S–CoM (apparent K m=250 M) of approximately 0.6 mol·min-1·mg protein-1. For activity the presence of Ti(III) citrate (apparent K m=15 M) and of ATP (apparent K m=30 M) were required in catalytic amounts. Ti(III) could be substituted by reduced ferredoxin. ATP could not be substituted by AMP, CTP, GTP, S-adenosylmethionine, or by ATP analogues. The membrane fraction was methylated by CH3–H4MPT in the absence of H–S–CoM. This methylation was dependent on Ti(III) and ATP. The methylated membrane fraction catalyzed the methyltransfer from CH3–H4MPT to H–S–CoM in the absence of ATP and Ti(III). Demethylation in the presence of H–S–CoM also did not require Ti(III) or ATP. Based on these findings a mechanism for the methyltransfer reaction and for the activation of the enzyme is proposed.Abbreviations H4MPT tetrahydromethanopterin - CH3–H4MPT N 5-methyl-H4MPT - H–S–CoM 2-mercaptoethanesulfonate or coenzyme M - CH3–S–CoM 2(methylthio)ethanesulfonate or methylcoenzyme M - SDS-PAGE sodium dodecylsulfate polyacrylamide gel electrophoresis - DTT dithiothreitol - MOPS morpholinopropanesulfonate - CHAPS 3-[(3-cholamidopropyl)-dimethylammonio]-1-propane-sulfonate - 1 U = 1 mol/min  相似文献   

15.
Summary Diffusion coefficients for FITC-molecular probes in intercellular pores (D) and rate of molecular probe loss into the vacuole (k1) have been obtained for FITC molecular probes in staminal hairs ofSetcreasea purpurea. The kinetic curves of FITC-Gly, -Ala, -Leu,-Ser, -Thr, -Cys, -Met, -Tyr, -Asp, -Glu, -Asn, -Gln, -Lys, -His,-Arg, -(Asp)2, -(Glu)2, -(Lys)2, -(Asp)3, -(Glu)3, -(Gln)2, -(Gln)3, -(Gln)4, and carboxyfluorescein (group I probes) matched the curves calculated for simple diffusion through a chain of cells, while the majority of kinetic curves of FITC-Phe, and -Try (group II probes) did not. None of the kinetic curves for FITC-(Met)2 and -(His)2 (group III probes) matched. Average Ds for group I probes ranged from 0.77× 10–8cm2/s to 3.75× 10–8cm2/s and for group II probes were 0.50× 10–8cm2/s. A meaningful average D for group III probes could not be calculated. Average k1 for group I probes ranged from 1.62× 10–7/m2/s to 13.21× 10–7/m2/s, and for group II probes were 5.42 and 11.54× 10–7/m2/s. Average k1s for group III probes could not be calculated. Symplastic transport occurred by cell-to-cell diffusion for most of the probes (e.g., group I probes) but not always for some (e.g., group II probes) and never for others (group III probes). The rate of cell-to-cell diffusion and loss within the vacuole depended upon the molecule's specific structure, molecular weight and charge. We concluded that plasmodesmata select for molecules that are hydrophilic, small and have a charge of from — 2 to — 4, and against molecules that contain either Phe, Try, Met or His groups.Abbreviations CF carboxyfluorescein - D diffusion coefficients for FITC-molecular probes in intercellular pores - k1 rate of FITC-molecular probe loss  相似文献   

16.
Summary Cotton (Gossypium hirsutum L. var. DP 61) was grown at different temperatures during 12-h light periods, with either 1800–2000 mol photons m–2 s–1 (high photon flux density, PFD) or 1000–1100 mol m–2 s–1 (medium PFD) incident on the plants. Night temperature was 25°C in all experiments. Growth was less when leaf temperatures were below 30°C during illumination, the effect being greater in plants grown with high PFD (Winter and Königer 1991). Leaf pigment composition and the photon-use efficiency of photosynthesis were analysed to assess whether plants grown with high PFD and suboptimal temperatures experienced a higher degree of high irradiance stress during development than those grown with medium PFD. The chlorophyll content per unit area was 3–4 times less, and the content of total carotenoids about 2 times less, with the proportion of the three xanthophylls zeaxanthin + antheraxanthin + violaxanthin being greater in leaves grown at 20–21°C than in leaves grown at 33–34°C. In leaves from plants grown at 21°C and 1800–2000 mol photons m–2 s–1, zeaxanthin accounted for as much as 34% of total carotenoids in the middle of the photoperiod, the highest level recorded in this study. This finding is consistent with a protective role of zeaxanthin under conditions of excess light. At the lower temperatures, the photochemical efficiency of photosystem II, measured as the ratio of variable to maximum fluorescence yield (F V/F M) after 12-h dark adaptation, was 0.76 in medium PFD plants and 0.75 in high PFD plants compared with 0.83 and 0.79, respectively, at the higher temperatures. The photon-use efficiency of O2 evolution () based on absorbed light between 630 and 700nm, decreased with decrease in temperature from 0.102 to 0.07 under conditions of high PFD, but remained above 0.1 at medium PFD. Owing to compensatory reactions in these long-term growth experiments, sustained differences inF V/F M and were much less pronounced than the differences in chlorophyll content and dry matter, particularly in plants which had developed at high PFD and low temperature. In fact, in these plants, which exhibited pronounced photobleaching, a largely functional photosynthetic apparatus was still maintained in cells adjacent to the lower leaf surfaces. This was indicated by measurements of photon use efficiencies of photosynthetic O2 evolution with leaves illuminated first at the upper, and then at the lower surface.Abbreviations F O yield of dark level fluorescence - F M maximum yield of fluorescence, induced in a pulse of saturating light - F V yield of variable fluorescence (=F M-F o) - PFD photon flux density - iw photon use efficiency of O2 evolution based on white (400–700 nm) incident light - ir photon use efficiency based on red (630–700 nm) incident light - aw photon use efficiency based on white absorbed light - ar photon use efficiency based on red absorbed light  相似文献   

17.
The CO2 gas exchange rates of the Central European perennial understory plantAsarum europaeum L. were measured in late autumn (October 30 to November 30) in its natural habitat day and night.During these measurements the temperature ranged from 0 to 15°C and the absolute air humidity from 3 to 10 mg H2O·1–1. Temperature and absolute air humidity over these ranges did not affect CO2 net assimilation which was determined almost entirely by quantum flux density.CO2 net assimilation was light saturated at about 100 M·m–2·s–1 quantum flux density. The uptake rate at this point was 4.3 mg·dm–2·h–1. The compensation point occurred at approximately 1 M·m–2·s–1.  相似文献   

18.
The interaction of various radioligands with spinal opioid receptors has been characterized under variable experimental conditions. Binding to , , and sites was measured in all (cervical, thoracic, lumbar) segments. The apparent affinity constant (K) of [3H]Ethylketocyclazocine (EKC) was similar in Tris, 2.09 (±1.06)×108 M–1, and phosphate buffer, 2.16 (±0.02)×108 M–1, when its interaction with and sites was blocked. Without blocking ligands, EKC binding was resolved in two components:K 1=1.01 (±0.21)×109 M–1 andK 2=0.95 (±0.61)×107 M–1. Likewise, the binding of [D-Ala2, MePhe4, Gly(ol)5]enkephalin (DAGO) or [D-Ala2, D-Leu5]-enkephalin (DADLE) alone was represented by a 2-site model. By adjusting the radioligand and receptor concentration or by the addition of blocking ligands, binding was represented by a 1-site model for DAGO,K=4.35 (±1.41)×108 M–1, and DADLE,K=2.44 (±0.08)×108 M–1.The abbreviations used are DADLE [D-Ala2, D-Leu5]enkephalin - DAGO [D-Ala2, MePhe4, Gly(ol)5]enkephalin - EKC ethylketocyclazocine - DYN dynorphin (1–17)  相似文献   

19.
Kinetic data of ferrous iron oxidation by Thionacillus ferrooxidans were determined. The aim was to remove H2S (<0.5 ppm) from waste gas by a process proposed earlier. Kinetic data necessary for industrial scale-up were investigated in a chemostat airlift reactor (dilution rate 0.02–0.12 h–1; pH 1.3). Due to the low pH, ferric iron precipitation and wall growth could be avoided. The maximum ferrous iron oxidation rate of submersed bacteria was 0.77 g 1–1 h–1, the maximum specific growth rate about 0.12 h–1 and the yield coefficient was found to be 0.007 g g–1 Fe2+. The specific O2 demand of an exponentially growing, ironoxidizing batch culture was 1.33 mg O2 mg–1 biomass h–1. The results indicate that a pH of 1.3 has no negative influence on the kinetics of iron oxidation and growth. Correspondence to: W. Schäfer-Treffenfeldt  相似文献   

20.
Bernhard Penth  Josef Weigl 《Planta》1971,96(3):212-223
Summary Influx of anions (5x10–4 M Cl or SO4 2–) across the plasmalemma, ATP levels and CO2 fixation in Limnophila and Chara have been measured in a comparative study.In Limnophila, influx, ATP level and CO2 fixation were progressively reduced by increasing concentrations of carbonyl cyanide m-chlorophenylhydrazone (CCCP) in the light (4000 lux) as well as in the dark. In Chara, not only influx but also ATP levels were much less reduced in the light than in the dark.At 5x10–4 M external salt concentration the action of light or dark is to change active influx of anions. Thus this study provides strong evidence to support the view that active anion uptake is directly dependent on ATP rather than on electron flow or NADPH. The possible significance of differences in the photophosphorylation systems of various plants is stressed.  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