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.
Summary Cell density is a factor that affects the capacity of Cloudman S91 melanoma cells to respond to melanotropins in monolayer culture. Continuous exposure of melanoma cells to α-melanotropin or its potent analog [Nle4,D-Phe7]-α-MSH, resulted in maximal stimulation of tyrosinase after 2 d of treatment, but the magnitude of stimulation decreased thereafter despite the continued presence of the melanotropins. However, when melanoma cells continually exposed to melanotropins were subcultured to an initial low cell density and maintained in contact with α-MSH or [Nle4,D-Phe7]-α-MSH (long-term culture), tyrosinase activity was rapidly restored and greatly enhanced. Also, when cells were seeded at initial densities ranging from 0.2 to 3.2×106 cells/flask, and exposed for 24 h to 10−7 M α-MSH, only the cultures seeded at low densities (0.2 and 0.4×106 cells/flask) exhibited maximal tyrosinase activity during the 24 h exposure to the melanotropins. Therefore, tyrosinase activity was primarily affected by cell density rather than by the duration of time the cells were in culture or by continuous exposure to melanotropin. Other flasks of various cell densities were treated with 10−7 M α-MSH or [Nle4,D-Phe7]-α-MSH for 24 h, followed byremoval of the melanotropins from the culture medium. The magnitude and duration of theresidual stimulation of melanoma tyrosinase activity by melanotropins were also found to be dependent on the initial cell density. These results reveal that there is a limited range of optimal cell densities at which melanoma cells can respond to melanotropins and express increased tyrosinase activity.  相似文献   

2.
Abstract Three kinds of trisaccharides were prepared by digesting fucoidan from the brown alga Kjellmaniella crassifolia, with the extracellular enzymes of the marine bacterium Fucobacter marina. Their structures were determined as Δ4,5GlcpUA1-2(L-Fucp(3-O-sulfate)α1-3)D-Manp, Δ4,5GlcpUA1-2(L-Fucp(3-O-sulfate)α1-3)D-Manp(6-O-sulfate), and Δ4,5GlcpUA1-2(L-Fucp(2,4-O-disulfate)α1-3)D-Manp(6-O-sulfate), which indicated the existence of a novel polysaccharide in the fucoidan and a novel glycosidase in the extracellular enzymes. In order to determine the complete structure of the polysaccharide and the reaction mechanism of the glycosidase, the fucoidan was partially hydrolyzed to obtain glucuronomannan, which is the putative backbone of the polysaccharide, and its sugar sequence was determined as (-4-D-GlcpUAβ1-2D-Manpα1-)n, which disclosed that the main structure of the polysaccharide is (-4-D-GlcpUAβ1-2(L-Fucp(3-O-sulfate)α1-3)D-Manpα1-)n. Consequently, the glycosidase was deduced to be an endo-α-D-mannosidase that eliminatively cleaves the α-D-mannosyl linkage between D-Manp and D-GlcpUA residues in the polysaccharide and produces the above trisaccharides. The novel polysaccharide and glycosidase were tentatively named as sulfated fucoglucuronomannan (SFGM) and SFGM lyase, respectively.  相似文献   

3.
The influence of sub-optimal temperatures (T) on the microbial growth rate (μ) has been assessed by means of dimensionless variables: Tdim = [T−Tmin]/[Topt−Tmin] and μdim = μ/μopt. Tmin represents the temperature at which there is no growth, Topt the optimum temperature at which the growth rate, μopt, is maximum. Data sets, growth rate vs temperature, have been taken from the literature for 12 organisms (psychrotrophs, mesophiles and thermophiles). In order to compare these organisms, the power law function has been used: [μdim] = [Tdim]α. The parameters μopt and Topt are determined from direct readings whereas Tmin and αare estimated by means of a non-linear regression. An accurate estimation of Tmin is obtained providing low growth rate data are available. A wide range of optimal temperatures where the growth rate almost equals μopt prevents one from obtaining a narrow confidence interval forα. On the basis of the analysis hereafter developed, thermophiles are characterized by values of the power α less than mesophiles and psychrotrophs. Almost all of these values are significantly different from two, previously determined for Staphylococcus xylosus and widely used for predicting the microbial growth in foods. Received 15 May 1998/ Accepted in revised form 25 September 1998  相似文献   

4.
Summary A Monte Carlo simulation is proposed to study the dynamics of helper T-cells (N H) and viral (N V) populations in an immune response model relevant to HIV. Cellular states are binary variables and the interactions are described by logical expressions. Viral population shows a nonmonotonic growth before reaching a constant value while helper T-cells grow to a constant after a relaxation/reaction time. Initially, the population of helper cells grows with time with a power-law, N Ht β, before reaching the steady-state; the growth exponent β increases systematically (β ≈ 1 – 2) with the mutation rate (P mut≈0.1–0.4). The critical recovery time (t c) increases exponentially with the viral mutation, t cAe αP mut , with α=4.52±0.29 in low mutation regime and α=15.21±1.41 in high mutation regime. The equilibrium population of helper T-cell declines slowly with P mut and collapses at ∼ 0.40; the viral population exhibits a reverse trend, i.e., a slow increase before the burst around the same mutation regime.  相似文献   

5.
Phylogenetic relationships among reptiles were examined using previously published and newly determined hemoglobin sequences. Trees reconstructed from these sequences using maximum-parsimony, neighbor-joining, and maximum-likelihood algorithms were compared with a phylogenetic tree of Amniota, which was assembled on the basis of published morphological data. All analyses differentiated α chains into αA and αD types, which are present in all reptiles except crocodiles, where only αA chains are expressed. The occurrence of the αD chain in squamates (lizards and snakes only in this study) appears to be a general characteristic of these species. Lizards and snakes also express two types of β chains (βI and βII), while only one type of β chain is present in birds and crocodiles. Reconstructed hemoglobin trees for both α and β sequences did not yield the monophyletic Archosauria (i.e., crocodilians + birds) and Lepidosauria (i.e., Sphenodon+ squamates) groups defined by the morphology tree. This discrepancy, as well as some other poorly resolved nodes, might be due to substantial heterogeneity in evolutionary rates among single hemoglobin lineages. Estimation of branch lengths based on uncorrected amino acid substitutions and on distances corrected for multiple substitutions (PAM distances) revealed that relative rates for squamate αA and αD chains and crocodilian β chains are at least twice as high as those of the rest of the chains considered. In contrast to these rate inequalities between reptilian orders, little variation was found within squamates, which allowed determination of absolute evolutionary rates for this subset of hemoglobins. Rate estimates for hemoglobins of lizards and snakes yielded 1.7 (αA) and 3.3 (β) million years/PAM when calibrated with published divergence time vs. PAM distance correlates for several speciation events within snakes and for the squamate ↔ sphenodontid split. This suggests that hemoglobin chains of squamate reptiles evolved ∼3.5 (αA) or ∼1.7 times (β) faster than their mammalian equivalents. These data also were used to obtain a first estimate of some intrasquamate divergence times. Received: 15 September 1997 / Accepted: 4 February 1998  相似文献   

6.
The thermodynamic parameters for six p53 carboxy-terminus peptide fragments as determined by analytical ultracentrifugal analysis were compared over the experimental temperature range of 275–310 K to evaluate the Gibbs free energy change as a function of temperature, ΔG o (T), from 0 to 400 K using our general linear third-order fitting function, ΔG o (T) = α + βT 2 + γT 3. Data obtained at the typical experimental temperature range are not sufficient to accurately describe the variations observed in the oligomerization of these p53 fragments. It is necessary to determine a number of thermodynamic parameters, all of which can be precisely assessed using this general third-order linear fitting function. These are the heat of reaction, innate temperature-invariant enthalpy, compensatory temperatures and the thermodynamic molecular switch occurring at the thermal set point. This methodology can be used to distinguish the characteristic structure and stability of p53 carboxy-terminal fragments or other p53 mutants. It should be used for the thermodynamic characterization of any interacting biological system.  相似文献   

7.
The thermophilic fungus Humicola sp constitutively produces intracellular α-galactosidase (1.33 U mg−1 protein) within 48 h at 45°C in shaken flasks, when grown in a medium containing 7% wheat bran extract as a carbon source and 0.5% yeast extract as a nitrogen source. The enzyme has been purified to homogeneity by ultrafiltration, ethanol precipitation, DEAE cellulose and Sephacryl S-300 chromatography with a 124-fold increase in specific activity and 29.5% recovery. The molecular weight of the enzyme is 371.5 kDa by gel filtration on Sephacryl S-300 and 87.1 kDa by SDS-polyacrylamide gel electrophoresis. The enzyme has an optimum temperature of 65°C and an optimum pH of 5.0. Humicola α-galactosidase is a glycoprotein with 8.3% carbohydrate content and is acidic in nature with a pI of 4.0. The K m S for p-nitrophenyl-α-D-galactopyranoside, O-nitrophenyl-α-D-galactopyranoside, raffinose and stachyose are 0.279, 0.40, 1.45 and 1.42 mM respectively. The enzyme activity was strongly inhibited by Ag+ and Hg2+. D-Galactose inhibited α-galactosidase competitively and the inhibition constant (K i) for galactose was 11 mM. Received 28 January 1999/ Accepted in revised form 07 April 1999  相似文献   

8.
An extracellular raw-starch-digesting α-amylase was isolated from Geobacillus thermodenitrificans HRO10. The culture conditions for the production of α-amylase by G. thermodenitrificans HRO10 was optimized in 1.2–l bioreactor using full 24 and 32 factorial designs. From the optimal reaction conditions, a model (Y = − 594.206 − 0.178T2 − 8.448pH2 + 6.020TpH − 0.005T2pH2) was predicted, which was then used for α-amylase production. In the bioreactor studies, the enzyme yield under optimized conditions (pH 7.1, 49°C) was 30.20 U/ml, a 51% improvement over the results (19.97 U/ml) obtained when the traditional one-factor-at-a-time method was employed. This α-amylase does not require extraneous calcium ions for activity, which may be a commercially important observation.  相似文献   

9.
The influence of temperature (T) and water activity (a w) on the growth rate (μ) of seven moulds (Alternaria alternata, Aspergillus flavus, Cladosporium cladosporioides, Mucor racemosus, Penicillium chrysogenum, Rhizopus oryzae and Trichoderma harzianum) was assessed in suboptimal conditions. Firstly, the dependence of fungal growth on temperature, at a w 0.99, was modelled through an approach described previously for bacteria. A dimensionless growth rate variable: μ dimα=μ/μ optα depended on the following normalised temperature: T dim=(TT min)/(T optT min) according to a power function: μ dimα=[T dim] α , where α was an exponent to be estimated. Secondly, the same approach was used to describe the influence of a w on fungal growth, at the respective optimum temperatures for each mould. Similarly, μ dimβ=μ/μ optβ depended on the following normalised water activity: a wdim=(a wa wmin)/(a wopta wmin) according to a power function: μ dimβ=[a wdim]β. Results show: (i) for each mould, the α-value is significantly less than the β-value, confirming that water activity has a greater influence than temperature on fungal development; (ii) the α-values and the β-values depend on the mould; (iii) the α-value is less than 1 for the mesophilic mould A. flavus, whereas the other moulds are characterised by higher α-values ranging from 1.10 to 1.54; (iv) the mesophilic A. flavus exhibits a low β-value, 1.50, compared to the hydrophilic T. harzianum, β=2.44, while β-values are within the range (1.71–2.37) for the other moulds. Journal of Industrial Microbiology & Biotechnology (2002) 28, 311–315 DOI: 10.1038/sj/jim/7000248 Received 27 June 2001/ Accepted in revised form 04 February 2002  相似文献   

10.
The effects of the substrate conditions on the volumetric productivity of Lactobacillus helveticus at different cell densities up to 60 g l−1 in a continuous stirred-tank reactor with microfiltration to retain the biomass were investigated. At low dilution rates, D, the steady-state volumetric productivity, r p, gradually increased to a maximum at D = 1.2–1.5 h−1, because of reduced product inhibition. At higher D values, r p unexpectedly decreased, although the substrate conditions further improved. The maxima of r p at different cell densities coincided with a critical specific substrate utilization rate beyond which the cell metabolism seems to be controlled through a catabolic modulator factor, and r p decreases. Received: 8 September 1997 / Received last revision: 31 December 1997 / Accepted: 2 January 1998  相似文献   

11.
In many studies on the protein folding problem it is assumed that the internal rotational barriers about NCα and CαC backbone bonds in unfolded polypeptides are quite small, around 0.7 kcal/mol, of an order comparable to the energy of kT at normal temperature (where k is Boltzmann’s constant and T is the temperature in K) and hence that rotations about these bonds occur almost freely. Here it is highlighted that such consideration is an unfortunate mistake. Approximate values for the rotational barriers of NCα and CαC bonds are suggested from computations of U(f \phi , ψ) potential energy surface (PES) maps of a number of oligopeptides by a semiempirical method for conformational analysis. The proposed values are about 16 kcal/mol for NCα bonds and 6 kcal/mol for CαC bonds. The values of the same barriers estimated from some ab initio quantum-mechanical PES maps for several dipeptides available in literature are also highlighted.  相似文献   

12.
Cell walls of Bacillus subtilis VKM B-760 and VKM B-764 are characterized by heterogeneous composition of teichoic acids. Polymer I with structure -6)-β-D-Galp-(1→1)-sn-Gro-(3-P-, polymer II with structure -6)-α-D-Glcp-(1→1)-sn-Gro-(3-P-, and a small amount of unsubstituted 1,3-poly(glycerol phosphate) were detected in strain VKM B-760. Strain VKM B-764 contains an analogous set of teichoic acids, but a characteristic feature of polymer II is the presence of disubstituted glycerol residue with α-glucopyranose localization in the integral chain at C-1 hydroxyl and β-glucopyranose as a side branch at C-2 hydroxyl (polymer III): -6)-α-D-Glcp-(1→1)-[β-D-Glcp-(1→2)]-sn-Gro-(3-P-. The structures of polymer I in bacilli and polymer III in Gram-positive bacteria are described for the first time. Teichoic acids were studied by chemical methods and on the basis of combined analysis of one-dimensional 1H-, 13C-, and 31P-NMR spectra, homonuclear two-dimensional 1H/1H COSY, TOCSY, and ROESY, and heteronuclear two-dimensional 1H/13C gHSQC- and HMQC-TOCSY experiments. Simultaneous presence of several different structure teichoic acids in the bacillus cell walls as well as chemotaxonomical perspectives of the application of these polymers as species-specific markers for members of the Bacillus genus is discussed.  相似文献   

13.
V. A. Bouryi 《Neurophysiology》1998,30(4-5):301-304
Barium currents through ion channels formed by α1-subunit of L-type Ca2+ channel (I α1) were recorded from cultured chinese hamster ovary (CHO) cells. The cells were stably transfected with either a cardiac or a smooth muscle (SM) variant of α1-subunit. TheI α1 in both cases exhibited similar fast voltage-dependent activation kinetics and slow apparent inactivation kinetics. With 10 mM Ba2+ in the bath solution,I α1 was activated at potentials more positive than −40 mV, peaked between 0 and +10 mV, and reversed at about +50 mV. In addition to slow apparent inactivation of inward current, both subunits provided an extremely slow voltage-dependent inactivation at potentials more positive than −100 mV, with half-maximum inactivation at −43.4 mV for cardiac and −41.4 mV for SM α1-subunits. The onset of inactivation as well as recovery from this process were within a time range of minutes. The voltage dependence of steady-state inactivation could be fitted by the sum of two Boltzmann's equations with slope factors of about 12 mV and 5 mV. A less sloped component has its midpoints at −75.6 and −63.7 mV, and a steeper component has its midpoints at −42.8 and −37.7 mV for cardiac and SM α1-subunits, respectively. Relative contribution of the steeper component was higher in both subunits (0.86 and 0.66 for cardiac and SM subunits, respectively). For comparison, the inactivation curves for 5-sec-long conditioning prepulses could be fitted by single Boltzmann's distribution with a 20 mV more positive midpoint and a slope factor of about 13 mV. In contrast to the steady-state inactivation curves, they showed considerable overlap with the steady-state activation curve. Our results reflect functional consequences of known sequence differences between α1-subunits of the cardiac and SM L-type Ca2+ channels and could be used in structural modeling of Ca2+ channel gating. In addition, they show that depolarization-induced window current has a transient nature and decays with the development of extremely slow inactivation. This is the first demonstration that slow inactivation of the L-type Ca2+ channel is an intrinsic property of its α1-subunits.  相似文献   

14.
For precise boundary conditions of biological relevance, it is proved that the steadily propagating plane-wave solution to the Fisher equation requires the unique (eigenvalue) velocity of advance 2(Df)1/2, whereD is the diffusivity of the mutant species andf is the frequency of selection in favor of the mutant. This rigorous result shows that a so-called “wrong equation”, i.e. one which differs from Fisher's by a term that is seemingly inconsequential for certain initial conditions, cannot be employed readily to obtain approximate solutions to Fisher's, for the two equations will often have qualitatively different manifolds of exact solutions. It is noted that the Fisher equation itself may be inappropriate in certain biological contexts owing to the manifest instability of the lowerconcentration uniform equilibrium state (UES). Depicting the persistence of a mutantdeficient spatial pocket, an exact steady-state solution to the Fisher equation is presented. As an alternative and perhaps more faithful model equation for the propagation of certain species properties through a homogeneous population, we consider a reaction-diffusion equation that features a cubic-polynomial rate expression in the species concentration, with two stable UES and one intermediate unstable UES. This equation admits a remarkably simple exact analytical solution to the steadily propagating plane-wave eigenvalue problem. In the latter solution, the sign of the eigenvelocity is such that the wave propagates to yield the “preferred” stable UES (namely, the one further removed from the unstable intermediate UES) at all spatial points ast→∞. The cubic-polynomial equation also admits an exact steady-state solution for a mutant-deficient or mutant-isolated spatial pocket. Finally, the perpetuating growth of a mutant population from an arbitrary localized initial distribution, a mathematical problem analogous to that for ignition in laminar flame theory, is studied by applying differential inequality analysis, and rigorous sufficient conditions for extinction are derived here.  相似文献   

15.
Small-angle neutron scattering (SANS) on the unilamellar vesicle (ULV) populations (diameter 500 and 1,000 Å) in D2O was used to characterize lipid vesicles from dimyristoylphosphatidylcholine (DMPC) at three phases: gel Lβ′, ripple Pβ′ and liquid Lα. Parameters of vesicle populations and internal structure of the DMPC bilayer were characterized on the basis of the separated form factor (SFF) model. Vesicle shape changes from nearly spherical in the Lα phase to elliptical in the Pβ′ and Lβ′ phases. This is true for vesicles prepared via extrusion through pores with the diameter 500 Å. Parameters of the internal bilayer structure (thickness of the membrane and the hydrophobic core, hydration and the surface area of the lipid molecule) were determined on the basis of the hydrophobic–hydrophilic (HH) approximation of neutron scattering length density across the bilayer ρ(x) and of the step function (SF) approximation of ρ(x). DMPC membrane thickness in the Lα phase (T=30°C) demonstrates a dependence on the membrane curvature for extruded vesicles. Prepared via extrusion through 500 Å diameter pores, vesicle population in the Lα phase has the following characteristics: average value of minor semi-axis 266±2 Å, ellipse eccentricity 1.11±0.02, polydispersity 26%, thickness of the membrane 48.9±0.2 Å and of the hydrophobic core 19.9±0.4 Å, surface area 60.7±0.5 Å2 and number of water molecules 12.8±0.3 per DMPC molecule. Vesicles prepared via extrusion through pores with the diameter 1,000 Å have polydispersity of 48% and membrane thickness of 45.5±0.6 Å in the Lα phase. SF approximation was used to describe the DMPC membrane structure in Lβ′ (T=10°C) and Pβ′ (T=20°C) phases. Extruded DMPC vesicles in D2O have membrane thickness of 49.6±0.5 Å in the Lβ′ phase and 48.3±0.6 Å in the Pβ′ phase. The dependence of the DMPC membrane thickness on temperature was restored from the SANS experiment.  相似文献   

16.
 Fourier transform infrared (FTIR) spectroscopy is used to compare the thermally induced conformational changes in horse, bovine and tuna ferricytochromes c in 50 mM phosphate/0.2 M KCl. Thermal titration in D2O at pD 7.0 of the amide II intensity of the buried peptide NH protons reveals tertiary structural transitions at 54  °C in horse and at 57  °C in bovine c. These transitions, which occur well before loss of secondary structure, are associated with the alkaline isomerization involving Met80 heme-ligand exchange. In tuna c, the amide-II-monitored alkaline isomerization occurs at 35  °C, followed by a second amide II transition at 50  °C revealing a hitherto unreported conformational change in this cytochrome. Amide II transitions at 50  °C (tuna) and 54  °C (horse) are also observed during the thermal titration of the CN-ligated cytochromes (where CN displaces the Met80 ligand), but a well-defined 35  °C amide II transition is absent from the titration curve of the CNadduct of tuna c. The different mechanisms suggested by the FTIR data for the alkaline isomerization of tuna and the mammalian cytochromes c are discussed. After the alkaline isomerization, loss of secondary structure and protein aggregation occur within a 5  °C range with T m values at 74  °C (bovine c), 70  °C (horse c) and 65  °C (tuna c), as monitored by changes in the amide I′ bands. The FTIR spectra were also used to compare the secondary structures of the ferricytochromes c at 25  °C. Curve fitting of the amide I (H2O) and amide I′ (D2O) bands reveals essentially identical secondary structure in horse and bovine c, whereas splitting of the α-helical absorption of tuna c indicates the presence of less-stable helical structures. CN adduct formation results in no FTIR-detectable changes in the secondary structures of either tuna or horse c, indicating that Met80 ligation does not influence the secondary structural elements in these cytochromes. The data provided here demonstrate for the first time that the selective thermal titration of the amide II intensity of buried peptide NH protons in D2O is a powerful tool in protein conformational analysis. Received: 1 April 1999 / Accepted: 24 August 1999  相似文献   

17.
By the use of the Immobiline technique at pH ranges 7.0–7.6 and 6.9–7.9, 16 different hemoglobin (Hb) phenotypes were observed in 61 English Saanen goats. They are explained in this breed by a genetic theory of five β-globin genes (A 4,A 6,A 8,E, andD) and two closely linked α-globin loci (′α and ″α) of which the ″α has a variant allele, provisionally called ″α X . Family data together with observed and expected Hb frequencies were in agreement with the genetic theory. Among six Barbary sheep there were three Hb phenotypes explained by the occurrence of the β-chain allelesB andC na.  相似文献   

18.
The protein α-globulin fromSesamum indicum L. has been characterised for its size and shape using αarious chemical, physico-chemical and hydrodynamic properties. The protein has an S20,w 0 of 12.8, D20,w °f 4.9 × 10-7 cm2/sec and a partial specific αolume of 0.725 ml/g in the natiαe state. The intrinsic αiscosity of the protein was determined to be 3 0 ml/g indicating it to be globular in shape. The molecular weight of the protein as determined by αarious approaches in analytical ultracentrifugation αaries from 2.6–2.74 × 105. The molecular weight from sedimentation equilibrium yields a αalue of 2.74 × 105 in the natiαe state and a αalue of 19000 in the dissociated and denatured state in 6 M guanidine hydrochloride. The eαaluation of frictional ratios using Stokes radius and results from electron microscopy confirms the protein to be globular in shape. The protein consists of at least 12–14 subunits. The eαaluation of hydrophobic parameters and energetics of interaction of subunits indicate that the protein is stabilized predominantly by hydrophobic interactions.  相似文献   

19.
A flow cytometry analysis and in vitro enzyme activity study is carried out on the methylotrophic yeast, Hansenula polymorpha, during both (a) batch growth and (b) continuous cultures subjected to single perturbations in either system dilution rate or influent carbon substrate composition. Flow cytometry of yeasts growing diauxically on a glucose: methanol mixture during exponential growth, exhibit DNA and RNA distributions indicative of the S-synthesis-phase of the cell cycle. Cells at the stationary growth stage exhibit DNA and RNA distributions that indicate one portion of the population in the G 0/G1 resting phase and another in the M-mitosis-phase.Yeast cells grown at a steady-state of D=0.2 h1, then shifted to D=0.35 h–1, at a constant influent substrate mixture, are also examined with both flow cytometry and in vitro enzyme assays. Distributions of DNA, RNA, and total protein at either steady state and during the shift between dilution rates did not resemble any observed in batch culture. Flow cytometry indicates significant changes in cell composition within 20 min of the imposed dilution rate shift. In vitro enzyme assays show a response time in decreasing methanol oxidase activity of 2.5–3 h upon a dilution rate shift-up, while hexokinase activity increases to its steady-state level in less than 3 h. Similar cell compositional changes are reported for shifts in influent substrate methanol: glucose ratio at a constant dilution rate of D=0.35 h –1. Results suggest that an unsteady-state regime, oscillating between conditions that promote maximum enzyme activity of either glucose- or methanol-metabolizing enzymes, may allow simultaneous enhanced time-averaged production of both sets of enzymes.  相似文献   

20.
Rapid light-response curves (RLC) of variable chlorophyll fluorescence were measured on estuarine benthic microalgae with the purpose of characterising its response to changes in ambient light, and of investigating the relationship to steady-state light-response curves (LC). The response of RLCs to changes in ambient light (E, defined as the irradiance level to which a sample is acclimated to prior to the start of the RLC) was characterised by constructing light-response curves for the RLC parameters α RLC, the initial slope, ETRm,RLC, the maximum relative electron transport rate, and E k,RLC, the light-saturation parameter. Measurements were carried out on diatom-dominated suspensions of benthic microalgae and RLC and LC parameters were compared for a wide range of ambient light conditions, time of day, season and sample taxonomic composition. The photoresponse of RLC parameters was typically bi-phasic, consisting of an initial increase of all parameters under low ambient light (E < 21–181 μmol m−2 s−1), and of a phase during which α RLC decreased significantly with E, and the increase of ETRm,RLC and E k,RLC was attenuated. The relationship between RLC and LC parameters was dependent on ambient irradiance, with significant correlations being found between α RLC and α, and between ETRm,RLC and ETRm, for samples acclimated to low and to high ambient irradiances, respectively. The decline of α RLC under high light (Δα RLC) was strongly correlated (P < 0.001 in all cases) with the level of non-photochemical quenching (NPQ) measured before each RLC. These results indicate the possibility of using RLCs to characterise the steady-state photoacclimation status of a sample, by estimating the LC parameter E k, and to trace short-term changes in NPQ levels without dark incubation.  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