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.
A compromise between carbon assimilation and structure investment at the leaf level is broadly accepted, yet the relationship between net assimilation per area (An) and leaf mass per area has been elusive. We propose bulk modulus of elasticity (ε) as a suitable parameter to reflect both leaf structure and function, and an inverse relationship between ε and An and mesophyll conductance (gm) is postulated. Using data for An, gm and ε from previous studies and new measurements on a set of 20 species covering all major growth forms, a negative relationship between An or gm and ε was observed. High ε was also related to low leaf capacitance and higher diffusive limitations to photosynthesis. In conclusion, ε emerges as a key trait linked with photosynthetic capacity across vascular plants, and its relationship with gm suggests the existence of a common mechanistic basis, probably involving a key role of cell walls.  相似文献   

2.
The in vivo reduction of ketone I (1,2-3H-dihydro(3,2,1-kl)pyridophenothiazine-3-one) by Rhizopus arrhizus Fisher is due to a NADPH-dependent alcohol dehydrogenase. This cytosolic enzyme displays a narrow specificity for ketone I, its pH optimum being pH 8. Partially purified alcohol dehydrogenase has a good affinity for ketone I (Km = 68 μM).  相似文献   

3.
4.
The ε subunits of several bacterial F1-ATPases bind ATP. ATP binding to the ε subunit has been shown to be involved in the regulation of F1-ATPase from thermophilic Bacillus sp. PS3 (TF1). We previously reported that the dissociation constant for ATP of wild-type ε subunit of TF1 at 25 °C is 4.3 μM by measuring changes in the fluorescence of the dye attached to the ε subunit (Kato, S. et al. (2007) J. Biol. Chem. 282, 37618). However, we have recently noticed that this varies with the dye used. In this report, to determine the affinity for ATP under label-free conditions, we have measured the competitive displacement of 2′(3′)-O-N′-methylaniloyl-aminoadenosine-5′-triphosphate (Mant-ATP), a fluorescent analog of ATP, by ATP. The dissociation constant for ATP of wild-type ε subunit of TF1 at 25 °C was determined to be 0.29 μM, which is one order of magnitude higher affinity than previously reported values.  相似文献   

5.
Acetoacetate decarboxylase is a valuable tool for clinical analysis of ketone bodies in human plasma. After screening many microorganisms, we found Bacillus polymyxa A-57 to be a new enzyme source with a good yield of acetoacetate decarboxylase. After purifying the intracellular enzyme by three-stage column chromatography, we identified it as a single protein by SDS-polyacrylamide gel electrophoresis. The enzyme had a molecular weight of approximately 280,000 and consisted of 10 (±2) identical subunits. It had an optimum pH of 5.9 and was stable up to about 60°C for 30min. The apparent Km and Vmax values for lithium acetoacetate were 0.94 mm and 296 μmiol/min/mg, respectively. In addition, the decarboxylase activity was found in the broth. After purification, we found that it was due to an active peptide which we named A-57-9, which we identified as the antibiotic polymyxin M1. However, the Vmax value (0.25 μmol/min/mg) of the peptide was very much lower and the Km value (400 mm) was higher than those of real acetoacetate decarboxylase.  相似文献   

6.
Ubiquitous Pseudomonads have great potential to influence the speciation and mobility of actinides in the environment. This study explores the unknown interaction between curium(III) and cell-suspensions of Pseudomonas fluorescens (CCUG 32456) isolated from the Äspö site, Sweden. The interaction between curium(III) and P. fluorescens cells was studied at trace curium(III) concentrations (0.3 μM) using time-resolved laser-induced fluorescence spectroscopy. Extraction studies have shown that the biosorption of curium(III) is a reversible process. Two Cm3+?P. fluorescens (CCUG 32456) species were identified, R?O?PO3H?Cm2+ and R?COO?Cm2+, having emission maxima at 599.6 and 601.9 nm, respectively. The corresponding surface complexation constants were determined to be log β111 = 12.7 ± 0.6 and log β110 = 6.1 ± 0.5, respectively.  相似文献   

7.
Glutamate dehydrogenase in Acanthamoeba castellanii is an NAD-dependent cytosolic enzyme. This is similar to glutamate dehydrogenases in Phycomycetes, but very different from the dual coenzyme-specific enzymes located in mitochondria in animals and in mitochondria and chloroplasts in higher plants. Pyrroline-5-carboxylate (P-5-C) reductase occurs also in the cytoplasm in A. castellanii and has very high affinities for L-P-5-C (Km= 12 μM) and NADH (Km= 15 μM). In contrast, ornithine aminotransferase and proline oxidase are mitochondrial enzymes. No proline-inhibited γ-glutamyl kinase was detected while an active glutamine synthetase was found in the cytosolic compartment. Evidence for a mitochondrial transport system for L-proline was obtained. Two possible pathways for proline biosynthesis in A. castellanii are discussed based on information obtained about activities and subcellular compartmentation of enzymes.  相似文献   

8.
Determination of DNA base compositions from melting profiles in dilute buffers   总被引:14,自引:0,他引:14  
Equations were determined for the dependency of the melting temperature (Tm) of DNA upon the logarithm of the sodium ion concentration, for four DNA samples of widely different base compositions (θGC). The slopes of these Tm versus log M equations wore found to decrease with increasing θG Cof the samples. An empirical equation relating Tm, log M (Na+) and θG C was derived, which also accounts for differences in Tm versus log M slopes. Data from the literature for some synthetic polynucleotides and for the crab(Cancer pagarus) satellite poly AT are discussed in relation to the above finding. The changes in Tm versus log M slopes with θG C are interpreted in terms of changes in the thermodynamic parameters ΔS and ΔH with base composition.  相似文献   

9.
Temperature-dependent uv absorption spectroscopy has been used to investigate the salt dependence of the order–disorder transition for the pH 4.2 rA8 double helix in 100% aqueous buffer and in a series of organic/aqueous mixed solvents. Melting temperature, Tm, data were obtained for the transitions in the different solvents by analysis of the uv melting curves. For the pure aqueous buffer solvent, the melting temperature was found to exhibit a reduced salt dependence (?tm/? log Na+) when compared to the corresponding polymer. This reduction is explained in terms of end effects and is shown to be consistent with the theoretical treatments of oligoelectrolyte transitions developed by Record and Lohman [Biopolymers, 17 , 159–166 (1978)]. In the mixed solvents, the salt dependence of the melting temperature (?tm/? log Na+) is shown to exhibit a linear dependence on the bulk dielectric constant of the medium for all of the hydroxyl-containing solvents studied. Significantly, N,N-dimethylformamide demonstrated different behavior.  相似文献   

10.
The inhibitory effects of 3-nitro-2,4,6-trihydroxybenzamide derivatives on human 5-lipoxygenase (5-LO), a key enzyme in arachidonic acid cascades, were examined using 5-LO produced by Escherichia coli. Some of the tested compounds inhibited the conversion of arachidonic acid to 5-hydroperoxy-6,8,11,14-eicosatetraenoic acid (5-HPETE), and in particular the N-phenylbutyl derivative was about 30 times more active (IC50 = 35 μm) than caffeic acid (IC50 = 1000 μm), a known selective 5-LO inhibitor.  相似文献   

11.
A methanol extract of neem oil indicated antifeedant activity at 200 μg/disc by a no-choice bioassay against Reticulitermes speratus Kolbe. The extract was purified by recycling HPLC to isolate 11 compounds of variable termite antifeedant potency. Deacetylgedunin was the most active compound (95% protective concentration or PC95 = 113.7 μg/disc). This was followed by salannin, gedunin, 17-hydroxyazadiradione, nimbandiol, azadiradione, deacetylsalannin, and deacetylnimbin, with PC95 estimates of 203.3, 218.4, 235.6, 245.4, 827.5, 1373.1, and 1581.2 μg/disc, respectively. Epoxyazadiradione, 17-epiazadiradione, and nimbin were not active (PC95 estimates were beyond the bioassay limits). It could be presumed, by comparing their structures and activities, that the furan ring, αβ-unsaturated ketone and hydroxyl group each played an important role in determining the antifeedant activity.  相似文献   

12.
M T Record 《Biopolymers》1967,5(10):993-1008
The theory developed in the previous paper to discuss changes in electrostatic free energies in polynucleotide order–disorder transitions is extended to cases where one or more of the participating species is titrated to some degree α. It is shown that, for any class of transition, the melting temperature Tm at constant pH is a linear function of the logarithm of the monovalent counterion concentration M, that at high salt the logarithm of the depression of the melting temperature by pH titration is proportional to the pH change, and that the stability of the ordered form as measured by its melting temperature at neutral pH, is a monotonic function of the quantity pHm – pK, where pHm and pK are the pH of melting and the monomer base pK, both measured under similar conditions of temperature and ionic strength. For the transition from double helix to coil, the dependences of Tm and dTm/d log M on pH are determined experimentally and compared with the qualitative predictions of the theory. It is found that dTm/dlog M, a measure of – ΔF?el (the negative of the electrostatic free energy change in the transition), decreases with increasing pH. In acid solution, where the coil is more extensively prolonated than the helix, the change in electrostatic free energy in the transition is larger than at neutral pH. Conversely, in alkali the electrostatic five energy change is smaller than at neutral pH. Hence (dTm/d log M)acid > (dTm/d log M neutral) > (dTm/d log M)alkali. At Suffeciently high pH, dTm/d log M is observed to become negative, indicating that the electrostatic free energy change is positive in the transition of this region. Date from the literature on the ionic strength dependence of the melting temperature for the acid helices of poly rA, poly rC, and poly dC are also considered from the standpoint of the theory.  相似文献   

13.
A quantitative light and electron microscope study of developing and degenerating mycorrhizal arbuscules of Glomus fasciculatum in Zea mays was carried out in order to estimate three parameters during the colonization cycle. These were: 1) Vv(f,c), the fraction of the host cell volume occupied by a volume of fungus; 2) Vv(cy,c), the fraction of the host cell volume occupied by host cytoplasm; 3) Sv(pr,c), the surface-area-to-volume ratio of the host protoplast to the whole host cell. Uninfected cortical cells had an Sv(pr,c) of 0.13 μm2/μm3. As the fungus penetrates the cell wall, the protoplast invaginates, causing a decrease in protoplast volume and an increase in protoplast Sv. The Sv(pr,c) of a cell containing a mature arbuscule is 1.275 μm2/μm3. Because of the shrinkage of the protoplast, the Sv of the protoplast to its own volume rather than the original cell volume is 2.55 μm2/μm3, or almost a 20-fold increase. Total cell size is unaffected. When the arbuscule is mature, the fungus occupies 42% of the cell, with 24% as 1-μm-diam branches, and 18% as trunk. Arbuscular branch formation progresses at a linear rate and is the most important factor in causing the increased host Sv. The correlation coefficient for Vv(br,c) the volume fraction for arbuscular branches, vs. Sv(pr,c) is r = 0.932 (P < 0.001). Degeneration of the arbuscule is marked by a rapid decrease in branches, host Sv, and host cytoplasm. The trunk develops and degenerates at a slower rate than the branches.  相似文献   

14.
This study presents a historical review, a meta‐analysis, and recommendations for users about weight–length relationships, condition factors and relative weight equations. The historical review traces the developments of the respective concepts. The meta‐analysis explores 3929 weight–length relationships of the type W = aLb for 1773 species of fishes. It shows that 82% of the variance in a plot of log a over b can be explained by allometric versus isometric growth patterns and by different body shapes of the respective species. Across species median b = 3.03 is significantly larger than 3.0, thus indicating a tendency towards slightly positive‐allometric growth (increase in relative body thickness or plumpness) in most fishes. The expected range of 2.5 < b < 3.5 is confirmed. Mean estimates of b outside this range are often based on only one or two weight–length relationships per species. However, true cases of strong allometric growth do exist and three examples are given. Within species, a plot of log a vs b can be used to detect outliers in weight–length relationships. An equation to calculate mean condition factors from weight–length relationships is given as Kmean = 100aLb?3. Relative weight Wrm = 100W/(amLbm) can be used for comparing the condition of individuals across populations, where am is the geometric mean of a and bm is the mean of b across all available weight–length relationships for a given species. Twelve recommendations for proper use and presentation of weight–length relationships, condition factors and relative weight are given.  相似文献   

15.
Marine phytoplankton and macroalgae acquire important resources, such as inorganic nitrogen, from the surrounding seawater by uptake across their entire surface area. Rates of ammonium and nitrate uptake per unit surface area were remarkably similar for both marine phytoplankton and macroalgae at low external concentrations. At an external concentration of 1 μM, the mean rate of nitrogen uptake was 10±2 nmol·cm?2·h?1 (n=36). There was a strong negative relationship between log surface area:volume (SA:V) quotient and log nitrogen content per cm2 of surface (slope=?0.77), but a positive relationship between log SA:V and log maximum specific growth rate (μmax; slope=0.46). There was a strong negative relationship between log SA:V and log measured rate of ammonium assimilation per cm2 of surface, but the slope (?0.49) was steeper than that required to sustain μmax (?0.31). Calculated rates of ammonium assimilation required to sustain growth rates measured in natural populations were similar for both marine phytoplankton and macroalgae with an overall mean of 6.2±1.4 nmol·cm?2·h?1 (n=15). These values were similar to maximum rates of ammonium assimilation in phytoplankton with high SA:V, but the values for algae with low SA:V were substantially less than the maximum rate of ammonium assimilation. This suggests that the growth rates of both marine phytoplankton and macroalgae in nature are often constrained by rates of uptake and assimilation of nutrients per cm2 surface area.  相似文献   

16.
B Y Tong  S J Battersby 《Biopolymers》1979,18(8):1917-1936
In this paper we analyze theoretically the observable details of the differential melting curves (DMC) and the denaturation maps (DM) of a DNA. With the help of a mathematical model, we explore their implications, their relation with each other and with the genetic map of the molecule, and discuss possible future applications. ?X174 is used as the example, since its sequence and genetic map are available. We find that each gene section of ?X174 has a characteristic DMC. A reconstruction scheme to get the DMC of a whole piece from those of its constituent genes is shown to be fairly successuful. The relations between the melting curve and the denaturation maps are clarified. We observe that nearly always, the beginning and end of a gene melt at lower temperatures. The sharp features in the DM indicate that despite the long-range cooperative interactions, the DM do reflect the local sequence effect. Denaturation maps (theoretical) of ?X174 and SV40 are presented. From available data of other authors, we estimate that the dependence of the melting temperature tm on GC, the fraction of (G+C)-content, and on x, the ionic concentration in fractions of the standard saline citrate solution, can be expressed as tm(x, GC) = -5.2 (log x)GC + 18.4 log x + 41.0GC + 69.4. The first two coefficients are less certain.  相似文献   

17.
Interaction between polylysine and DNA's of varied G + C contents was studied using thermal denaturation and circular dichroism (CD). For each complex there is one melting band at a lower temperature tm, corresponding to the helix–coil transition of free base pairs, and another band at a higher temperature tm, corresponding to the transition of polylysine-bound base pairs. For free base pairs, with natural DNA's and poly(dA-dT) a linear relation is observed between the tm and the G + C content of the particular DNA used. This is not true with poly(dG)·poly(dC), which has a tm about 20°C lower than the extrapolated value for DNA of 100% G + C. For polylysine-bound base pairs, a linear relation is also observed between the tm and the G + C content of natural DNA's but neither poly(dA-dT) nor poly(dG)·poly(dC) complexes follow this relationship. The dependence of melting temperature on composition, expressed as dtm/dXG·C, where XG·C is the fraction of G·C pairs, is 60°C for free base pairs and only 21°C for polylysine-bound base pairs. This reduction in compositional dependence of Tm is similar to that observed for pure DNA in high ionic strength. Although the tm of polylysine-poly(dA-dT) is 9°C lower than the extrapolated value for 0% G + C in EDTA buffer, it is independent of ionic strength in the medium and is equal to the tm0 extrapolated from the linear plot of tm against log Na+. There is also a noticeable similarity in the CD spectra of polylysine· and polyarginine·DNA complexes, except for complexes with poly(dA-dT). The calculated CD spectrum of polylysine-bound poly(dA-dT) is substantially different from that of polyarginine-bound poly(dA-dT).  相似文献   

18.

Scenedesmus is a genus of microalgae employed for several industrial uses. Industrial cultivations are performed in open ponds or in closed photobioreactors (PBRs). In the last years, a novel type of PBR based on immobilized microalgae has been developed termed porous substrate photobioreactors (PSBR) to achieve significant higher biomass density during cultivation in comparison to classical PBRs. This work presents a study of the growth of Scenedesmus vacuolatus in a Twin Layer System PSBR at different light intensities (600 μmol photons m−2 s−1 or 1000 μmol photons m−2 s−1), different types and concentrations of the nitrogen sources (nitrate or urea), and at two CO2 levels in the gas phase (2% or 0.04% v/v). The microalgal growth was followed by monitoring the attached biomass density as dry weight, the specific growth rate and pigment accumulation. The highest productivity (29 g m−2 d−1) was observed at a light intensity of 600 μmol photons m−2 s−1 and 2% CO2. The types and concentrations of nitrogen sources did not influence the biomass productivity. Instead, the higher light intensity of 1000 μmol photons m−2 s−1 and an ambient CO2 concentration (0.04%) resulted in a significant decrease of productivity to 18 and 10–12 g m−2 d−1, respectively. When compared to the performance of similar cultivation systems (15–30 g m−2 d−1), these results indicate that the Twin Layer cultivation System is a competitive technique for intensified microalgal cultivation in terms of productivity and, at the same time, biomass density.

  相似文献   

19.
Lipids of Algae     
The carotenoid pigments prepared from acetone extracts of chlorella were separated into epiphasic and hypophasic fractions by partition between petroleum ether and 90% methanol. Each fraction was subjected to column chromatography, using aluminium oxide, magnesium oxide, calcium hydroxide or calcium carbonate as adsorbent. The absorption maxima of the separated pigments in hexane and in carbon disulfide were compared with those of the known pigments. Some of the separated pigments were identified as those previously known which follow: α-carotene, β-carotene, rhodoxanthin, sarcinaxanthin, lutein and neoxanthin. Two unknown pigments with absorption maxima not yet reported were separated. The first showed absorption maxima at 478 mμ in hexane and 518 mμ in carbon disulfide, and the second at 383, 402 and 425 mμ in hexane and 428 and 450 mμ in carbon disulfide.  相似文献   

20.
H J Li  B Brand  A Rotter  C Chang  M Weiskopf 《Biopolymers》1974,13(8):1681-1697
Thermal denaturation of direct-mixed and reconstituted polylysine–DNA complexes in 2.5 × 10?4 M EDTA, pH 8.0 and various concentrations of NaCl has been studied. For both complexes, increasing ionic strength of the solution raises Tm, the melting temperature of free base pairs. The linear dependence of Tm on log Na+ indicates that the concept of electrostatic shielding on phosphate lattice of an infinitely long pure DNA by Na+ can be applied to short free DNA segments in a nucleoprotein. For a direct-mixed polylysine–DNA complex, the melting temperature of bound base pairs Tm′ remains constant at various ionic strengths. On the other hand, the Tm′ in a reconstituted polylysine–DNA complex is shifted to lower temperature at higher ionic strength. This phenomenon occurs for reconstituted complex with long polylysine of one thousand residues or short polylysine of one hundred residues. It is shown that such a decrease of Tm′ is not due to a reduction of coupling melting between free and bound regions in a complex when the ionic strength is raised. It is also not due to intermolecular or intramolecular change from a reconstituted to a direct-mixed complex. It is suggested that this phenomenon is due to structural change on polylysine-bound regions by ionic strength. It is suggested further that Na+ may replace water molecules and bind polylysine-bound regions in a reconstituted complex. Such a dehydration effect destabilizes these regions and lowers Tm′. This explanation is supported by circular dichroism (CD) results.  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