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 46 毫秒
1.
Reconstitution of the apoprotein of the molybdoenzyme nitrate reductase in extracts of the Neurospora crassa mutant nit-1 with molybdenum cofactor released by denaturation of purified molybdoenzymes is efficient in the absence of exogenous MoO42? under defined conditions. Evidence is presented that this molybdate-independent reconstitution is due to transfer of intact Mo cofactor, a complex of Mo and molybdopterin (MPT), the organic constituent of the cofactor. This complex can be separated from denatured protein by gel filtration, and from excess MoO42? by reverse-phase HPLC. Sulfite oxidase, native xanthine dehydrogenase, and cyanolyzed xanthine dehydrogenase are equipotent Mo cofactor donors. Other well-studied inactive forms of xanthine dehydrogenase are also shown to be good cofactor sources. Using xanthine dehydrogenase specifically radiolabeled in the cyanolyzable sulfur, it is shown that this terminal ligand of Mo is rapidly removed from Mo cofactor under the conditions used for reconstitution.  相似文献   

2.
Two ferredoxins from nitrogen-fixing cells of the phototrophic bacterium Rhodopseudomonas capsulata, strain B10, are purified to a homogeneous state and characterized. The molecular mass of ferredoxin I is about 12 kDa and that of ferredoxin II, 18 kDa. Ferredoxin I contains 8 Fe2+ and 8 S2?; ferredoxin II has 4 Fe2+ and 4 S2? per molecule. The redox potential of ferredoxin I is about ?270 mV and that of ferredoxin II ?419 mV. Ferredoxin I is more labile to the action of O2, O?2, H2O2 and heating. The ferredoxins are also different in their absorption and EPR spectra, amino acid composition and electron-transfer activity to Rps. capsulata nitrogenase: both C2H2 reduction and H2 evolution by Rps. capsulata nitrogenase proceed faster in the presence of ferredoxin I than in case of ferredoxin II. Synthesis of ferredoxin I takes place only in Rps. capsulata nitrogen-fixing cells grown in light under anaerobic conditions whereas ferredoxin II formation does not depend on the source of nitrogen or the growth medium, though the amount of ferredoxin II varies with the growth conditions. Its highest level has been found in the cells grown in lactate-limited medium in the presence of CO2 and light or in the presence of glutamate in darkness under anaerobic conditions.  相似文献   

3.
Six new dinuclear complexes, derived from cis-[Co(H2O)2(NH3)4]3+, cis-[Co(H2O)2(en)2]3+ and [M(CN)42? (M = Ni, Pd, Pt) were prepared and characterized by means of chemical analysis, electronic and IR measurements. The influence of the pH on the rate of the reaction was studied for the two derivatives of [Pd(CN)4]2?, showing that the best conditions to obtain the dinuclear compounds are at pH near 6, where the predominant species are cis-[Co(OH)(H2O)(amine)2]2+. The [Pt(CN)4]2? derivatives show PtPt interactions both in the solid state and in solution.  相似文献   

4.
The binding of[Co(CN)6]3?, and that of[Fe(CN)6]3? and [Ru(CN)6]4? using a competitive method, to horse cytochrome c has been studied by 59 Co NMR spectroscopy. At I = 0.07 M, without added salt and in 2H2O at ph* 7.3 (measured in 2H2O) and 25°C, there are at least two binding sites on ferricytochrome c and ferrocytochrome c for [Co(CN)6]3?. Association constants were determined to be 2.0 ± 0.6 × 103M?1 and 1.5 ± 0.5 × 102M?1 respectively. with no effect of the oxidation state of the cytochrome. At higher ionic strength (I = 0.12 M adjusted with KCl the binding markedly decreased, and, although it was not possible to determine the precise binding stoichiometry and magnitude of association constants, it is clear that the association constants are ≤ 1.5 × 10tM?1 The binding of [Ru(CN)6]4? at I = 0.07, without added salt and in 2H2O at pH 1.3 and 23°C, was not precisely defined, but its binding strength relative to that of [Fe(CN)6]3? was determined. Extrapolating this to I = 0.12 (KCl) suggests that under these conditions the association constant for [Ru(CN)6]4? binding to ferricytochrome c is ≤ 3 × 102M?1.  相似文献   

5.
6.
Complexation in the H+-Si(OH)4-tropolone (HL) system was studied in 0.6 M (Na)Cl medium at 25° C. Speciation and formation constants were determined from potentiometric (glass electrode) and 29Si-NMR data. Experimental data cover the ranges 1.5 ? - log[H+] ? 8.4, 0.002 ? B ? 0.012 M, and 0 ? C ? 0.060 M (B and C stand for the total concentration of Si and tropolone, respectively). In acid solutions (-log[H+] ? 3) a hexacoordinated cationic complex, SiL3+, is formed with log K(Si(OH)4 + 3HL + H+ XXX SiL3+ + 4H2O) = 7.08 ± 0.03. Furthermore, the formation of a disilicic acid was established from 29Si-NMR data. The dimerization constant of Si(OH)4 was found to be 10 exp (1.2 ± 0.1). In model calculations the solubility of quartz and amorphous SiO2 in the presence of tropolone is demonstrated. Data were analyzed using the least-squares computer program LETAGROPVRID.  相似文献   

7.
Kinetic constants for SO42? transport by upper and lower rat ileum in vitro have been determined by computer fitting of rate vs concentration data obtained using the everted sac technique. MoO42? inhibition of this transport is competitive, and kinetic constants for the inhibition were similarly determined. Transport is also inhibited by the anions WO42?, S2O32? and SeO42?, in the order S2O32? > SeO42? ≧ MoO42? > WO42?. These anions have no effect on the transport of l-valine. Low SO42? transport rates were observed in sacs from animals fed a high-molybdenum diet. The significance of the results with respect to the problem of molybdate toxicity in animals is discussed, and related to the known protective effect of SO42?.  相似文献   

8.
Electrospray ionization mass spectrometry (ESI MS) has been conducted on the ammonium and alkali metal (A=Li+, Na+ and K+) dichromate systems. A large number of previously unknown polyoxochromate species have been characterized. Major series that have been identified include [Ax+1HxCrVIxO4x]+ (Li+, x=1-5; Na+, x=1-7; K+, x=1-4) and [A2x−1CrVIxO4x−1]+ (Li+, x=2, 3; Na+, x=2-4; K+, x=2, 3) in the alkali metal dichromate systems, and [HCrVIxO3x+1] (x=1-5) in the ammonium dichromate system. Several series also contain mixed oxidation state species, ranging from Cr(V) to Cr(II) in conjunction with Cr(VI), which is consistent with the ease of reduction of Cr(VI). Negative ion ESI MS spectra clearly demonstrate the existence of [HCrO4] as the most abundant ion at −20 V, suggesting that its existence in solution is not just hypothetical, as was previously thought. The polymerization units for the series observed include {AHCrO4}, {A2CrO4} and {CrO3}, with the latter prominent in the alkali metal systems. This presumably arises from the fragmentation of dichromate, A2Cr2O7→{A2CrO4}+{CrO3}. Moreover, the ESI MS of the dichromate compounds have illustrated that the preservation of tetrahedral stereochemistry is of paramount importance for these systems, which leads to only limited polymerization compared to the related molybdate and tungstate systems.  相似文献   

9.
1-adrenaline, ACTH and glucagon activate the adenylate cyclase of rat adipocytes by decreasing its S0.5(Mg2+) (concentration yielding 0.5 Vmax) from its basal value of 11.5 to 1.2, 0.3 and 1.8 mM and by increasing its Ki(ATP4?) from 0.03 to 0.25; 0.62 and 0.16 mM respectively. The kinetic properties of the enzyme are regulated by its state of saturation with ATP4? or Mg2+; its saturation with ATP4? and citrate3? suppressed its basal and hormone-dependent activities. The hormone-dependent decrease in Km and increase in Vmax of the enzyme occur when shifting from suboptimal low concentrations of hormone and Mg2+ to optimal conditions, i.e., high concentration of hormone and low concentration of Mg2+. The increase in the state of saturation of the enzyme with Mg2+ decreases the hormone-dependent effects on Vmax and results in identical values of Km (0.14 mM) for its basal and 1-adrenaline dependent activities. CaCl2 saturation curves at 5 mM ATP with either 5, 10 or 20 mM MgCl2 show that the substitution of 5 mM MgCl2 by 10 mM and 20 mM MgCl2 increased the Ki(Ca2+) of the enzyme from 0.19 to 0.49 and 0.94 mM but decreased its Ki(CaATP) from 0.42 to 0.19 and 0.14 mM respectively. Only when the concentration of MgCl2 exceeded that of ATP did 1-adrenaline and ACTH activate the enzyme by increasing its Ki(Ca2+), although only ACTH increased its Ki(CaATP). An increase in energy charge would decrease the intracellular concentrations of Mg2+ and Ca2+ because ATP4? has stronger binding constants for Mg2+ and Ca2+ than ADP3? and AMP2?. Hence, the reported properties of the enzyme suggests that changes in energy charge may allow for metabolic feedback control of the hormonal responsiveness of the Mg2+, Ca2+, ATP4? -sensitive adenylate cyclase.  相似文献   

10.
Various His-Pt(II) coordination compounds were prepared by reaction of K2PtCl4 or cis-[Pt(NH3)2Cl2](cis-DDP) with His and analyzed by 1H and 13C NMR spectroscopy, electrophoresis, and ion-exchange chromatography. His may be coordinated to Pt by the imidazol iminogroup and/or the α-aminogroup; the carboxy group remains always free. Both bidentate as well as monodentate ligands were identified. Cis-DDP reacts with His to give a mixture of compounds where all these possibilities are present: cis-diamine-(histidine-N,N-)Pt(II) and three different types of cis-diammine-bis(histidine). HCl trans cleavage of compounds with bidentate His ligands leads to a mixture of two compounds having His ligated to Pt by an amino or imin group. The methods applied are suitable for analyzing reactions of His with cis-DDP under model conditions similar to physiological conditions.  相似文献   

11.
About 68–86% of the cysteine synthase activity in leaf tissue of white clover (Trifolium repens) and peas (Pisum sativum cultivar Massey Gem) was associated with chloroplasts. The enzymes from white clover and peas were purified ca 66 and 12-fold respectively. For clover, the Km values determined by calorimetric and S2? ion electrode methods were: S2? 0.51 and 0.13 mM; O-acetylserine (OAS), 3.5 and 2.O mM respectively. The analogous values for the pea enzyme were: S2?, 0.24 and 0.06 mM; OAS, 3.1 and 0.24 mM. Both enzymes were inhibited by cystathionine and cysteine. Pretreatment with cysteine inactivated the enzyme, but addition of pyridoxal phosphate caused partial reactivation. Isolated pea chloroplasts (70–75 % intact) catalysed OAS-dependent assimilation of sulphide at a mean rate of 88 μmol/mg Chl/hr. About 85 % of the OAS-dependent sulphide assimilated was recovered as cysteine. The rates were unaffected by light and 2 μM DCMU. Sonicating the chloroplasts enhanced the rate by 1.3–2 fold. Cysteine synthase activity was associated with the chloroplast stroma. Similar results were obtained for clover chloroplasts except that both the intactness and the rates were lower.  相似文献   

12.
The oxidation enthalpy of reduced flavin mononucleotide at pH 7.0 in 0.2 m phosphate buffer has been studied by determining the heat associated with the reaction: FMNH2 + 2 Fe(CN)?36 ? FMN + 2 Fe(CN)?46 + 2 H+. (a) (The quinone, semiquinone, and hydroquinone forms of FMN are represented as FMN, FMNH, and FMNH2, respectively.) Calorimetric experiments were performed in a flow microcalorimeter which was modified to prevent sample contamination by oxygen. The enthalpy observed for reaction (a), after correction for dilution and buffer effects, was ?39.2 ± 0.4 kcal (mole FMNH2)?1 at 25 °C. The potential difference, ΔE′, developed by reaction (a) was determined potentiometrically and corresponded to a free energy change, ΔG′, of ?30.3 kcal (mole FMNH2)?1. The resulting entropy change, ΔS′, was thus calculated to be ?29.8 e.u. Reaction (a) was also studied at temperatures of 7 °C and 35.5 °C. ΔCp′ for the reaction was calculated as ?155 ± 18 cal deg?1 (mole FMNH2)?1 at 20 °C. ΔH′ for the reaction (b), FMNH2 ? FMN + H2, (b) was calculated as +14.2 ± 0.7 kcal mole?1 at 25 °C, relative to the enthalpy of the hydrogen electrode being identically equal to zero at all values of pH and temperature. The free energy at pH 7.0 for reaction (b), calculated from the potential was found to be ?9.7 kcal mole?1, which resulted in an entropy for reaction (b) of 80.2 e.u. A thermal titration of reaction (a) was used to calculate the thermodynamic parameters for the formation of semiquinone dimer according to the reaction FMNH2 + FMN ? (·FMNH)2. (c) The free energy, enthalpy, and entropy changes for reaction (c) were estimated to be ?6.1 kcal mole?1, ?7 kcal mole?1, and ?3 e.u., respectively.  相似文献   

13.
Thrips palmi Karny has recently been implicated in damage to eggplant fruits. This thrips causes cosmetic scars on and deforms fruits, thus lowering their market value. To investigate the relationship between the density of thrips and the resulting damage, experimental plots with initial release densities of 0, 20, 40, 70, and 100 adults thrips per plot were established under greenhouse conditions. Thrips density (for flower sampling = x 1; for sticky trap = x 2) was monitored using flower sampling and blue sticky trap counts, and was found to be directly related to the proportion of damaged-fruits (y) and the reduction in fruit yield: significant relationships were found for the flower samples (y = 1.2261x 1 ? 0.6232, r 2 = 0.8582) and for the trap catches (y = 11.667 ln(x 2) ? 9.5, r 2 = 0.8896). The proportion of damaged fruits that could be tolerated from an economic perspective, based on the cost of controlling the thrips population chemically and the market value of the fruit, was 6.67–11.76 %; this translated into economic thresholds of 1.05–1.5 adults and/or larvae of T. palmi per flower or 4.91–10.17 adults T. palmi per four-day sticky card count.  相似文献   

14.
Microalgae-based CO2 capture from flue gas is an attractive mitigation strategy in the cement industry. However, NO x and SO x components might be harmful to microalgae. We performed toxicity assays, under 2 % (v/v) CO2 and using nitrite, sulfite, or bisulfite salts, on an environmental isolate, identified as Desmodesmus abundans (The University of Texas at Austin (UTEX), no. 2976) and Scenedesmus sp. UTEX1589. Nitrite and sulfite did not inhibit growth at the tested concentrations (0–1,067 ppm (w/v) NO2 ? and 0–254 ppm (w/v) SO3 2?); however, bisulfite was toxic above 39 ppm. Non-toxic concentrations of both sulfur-based compounds stimulated growth, but significantly higher growth rates were only observed for HSO3 ?. Within a narrower range, NO x and SO x served as a sole nutrient source. Overall, biomass production and growth rates of the environmental isolate were greater. A novel strategy to buffer high concentrations of HSO3 ? (200 ppm) was developed by adding cement kiln dust (CKD), a byproduct and flue gas component. The results suggest that CKD also provided other beneficial growth components and that sulfur optimization of the culture medium significantly increased carbon assimilation, particularly in D. abundans. In additional simulations of typical flue gas conditions in a modern cement plant (320, 40, and 40 ppm (w/v) of NO2 ?, SO3 2?, and HSO3 ?, respectively, and 25 % (v/v) CO2), along with the incorporation of 300 ppm CDK, growth of D. abundans was supported. Although further studies are needed, direct utilization of flue gas might be possible with the environmental isolate, where NO x , SO x , and CKD are all beneficial components of the mitigation system.  相似文献   

15.
《Inorganica chimica acta》1986,123(4):193-196
The stoichiometry of the reaction between Cu(II) and MoS42− in neutral aqueous solution was observed to proceed with a 1.5:1 Cu:Mo ratio. The reaction results in the reduction of Cu(II) and the quantitative formation of an insoluble solid. The results contrast with an earlier report of a 1:1 stoichiometric ratio, this latter ratio was, however, observed with CuII(albumin) as reactant, in this case no precipitate was observed. The insoluble products were examined by a number of spectroscopic techniques and by X-ray power diffraction and elemental analysis. Two products were identified. Solid A is isostructural with the known (NH4)CuIMoVVIS4, i.e. has the composition MICuMoS4, MI= NH4+, Na+ or Et4N+. Solid B has the approximate composition CuMoS4Ox, x=2−3 and contains Cu(I) and Mo(V) centres. Formation of compound B therefore involves an unusual internal two-electron redox process. The reaction and products are of particular biological significance.  相似文献   

16.
The title compound has been synthesized and subjected to crystal structure analysis. Mr = 548.50, m.p. 108.1 °C (decom.), orthorhombic, Im2m,a = 7.006(2), b = 8.938(2), c = 13.619(2) Å V = 852.8(3) Å3, Z = 2, Dx = 2.136, Dm, (flotation in CCl4/CH2I2) = 2.128 g cm?3, λ(Mo-Kα) = 0.71069 Å, μ = 90.79 cm?1, F(000) = 519.89, T = 295 K, final RF = 0.036 and RG = 0.044 for 566 observed reflections. The discrete [UO2F4(H20)]2? anion has site symmetry m2m, its virtually linear uranyl moiety being surrounded by fluoro and aquo ligands occupying the vertices of a pentagon in the equatorial plane. Watet molecules serve to link the complex anions by hydrogen bonds into layers, between which the organic cations are accommodated.  相似文献   

17.
The observed equilibrium constants (Kobs) for the reactions of d-2-phosphoglycerate phosphatase, d-2-Phosphoglycerate3? + H2O → d-glycerate? + HPO42?; d-glycerate dehydrogenase (EC 1.1.1.29), d-Glycerate? + NAD+ → NADH + hydroxypyruvate? + H+; and l-serine:pyruvate aminotransferase (EC 2.6.1.51), Hydroxypyruvate? + l-H · alanine± → pyruvate? + l-H · serine±; have been determined, directly and indirectly, at 38 °C and under conditions of physiological ionic strength (0.25 m) and physiological ranges of pH and magnesium concentrations. From these observed constants and the acid dissociation and metal-binding constants of the substrates, an ionic equilibrium constant (K) also has been calculated for each reaction. The value of K for the d-2-phosphoglycerate phosphatase reaction is 4.00 × 103m [ΔG0 = ?21.4 kJ/mol (?5.12 kcal/mol)]([H20] = 1). Values of Kobs for this reaction at 38 °C, [K+] = 0.2 m, I = 0.25 M, and pH 7.0 include 3.39 × 103m (free [Mg2+] = 0), 3.23 × 103m (free [Mg2+] = 10?3m), and 2.32 × 103m (free [Mg2+] = 10?2m). The value of K for the d-glycerate dehydrogenase reaction has been determined to be 4.36 ± 0.13 × 10?13m (38 °C, I = 0.25 M) [ΔG0 = 73.6 kJ/mol (17.6 kcal/mol)]. This constant is relatively insensitive to free magnesium concentrations but is affected by changes in temperature [ΔH0 = 46.9 kJ/mol (11.2 kcal/mol)]. The value of K for the serine:pyruvate aminotransferase reaction is 5.41 ± 0.11 [ΔG0 = ?4.37 kJ/mol (?1.04 kcal/mol)] at 38 °C (I = 0.25 M) and shows a small temperature effect [ΔH0 = 16.3 kJ/ mol (3.9 kcal/mol)]. The constant showed no significant effect of ionic strength (0.06–1.0 m) and a response to the hydrogen ion concentration only above pH 8.5. The value of Kobs is 5.50 ± 0.11 at pH 7.0 (38 °C, [K+] = 0.2 m, [Mg2+] = 0, I = 0.25 M). The results have also allowed the value of K for the d-glycerate kinase reaction (EC 2.7.1.31), d-Glycerate? + ATP4? → d-2-phosphoglycerate3? + ADP3? + H+, to be calculated to be 32.5 m (38 °C, I = 0.25 M). Values for Kobs for this reaction under these conditions and at pH 7.0 include 236 (free [Mg2+] = 0) and 50.8 (free [Mg2+] = 10?3m).  相似文献   

18.
Due to the industrial development, drinking water is getting polluted by disposing several waste products of the industries. Hardness is one of the prominent impurities in drinking water which is mainly due to the presence of carbonate and bicarbonate ions (CO3 2? and HCO3 ?) in it. Here, we present the synthesis of the zinc oxide (ZnO) and polyaniline (PANI) nanocomposite for the detection and estimation of hardness of the drinking water. The chemical formula of such a nanocomposite is defined in terms of the fraction of polyaniline nanoparticles reinforced in ZnO matrix and is derived as ZnO(1???x)PANI x (0?≤?x?≤?0.9); x is the composition ratio. Silver and ZnO(1???x)PANI x layers are coated over the unclad core of the optical fiber so as to create the four layer system as that of Kretschmann configuration SPR structure. The working principle involves the change in dielectric constant of (ZnO(1???x)PANI x ) by CO3 2? or HCO3 ? ions in aqueous atmosphere. Due to the strong interaction of the sensing surface to the CO3 2? and HCO3 ? ions, a red shift in the SPR spectrum is observed in the concentration range 0–200 μg/l. The sensitivity of the sensor depends on the composition ratio of the nanocomposite and has been found to be maximum for the composition ratio lying in the range 0.45–0.60. This has been further confirmed in terms of the enhancement of the electric field density and found to be in agreement with the experimental value. The sensitivity of the sensor with optimum value of the composition ratio is 0.094 and 0.065 nm/(μg/l) for CO3 2? and HCO3 ?, respectively. The sensor is highly selective to CO3 2? and HCO3 ?. The sensor has advantages of online monitoring and remote sensing of water quality because the probe is fabricated over an optical fiber.  相似文献   

19.
This study was undertaken to define optimal conditions for exchange of 3H-R-1881 with endogenous hormone bound to androgen receptor (AR) sites in homogenates of rat ventral prostate (RVP) of mature animals. To minimize inactivation of AR binding sites under exchange conditions, extracellular proteases present in RVP were removed by mincing and washing tissue fragments in a Ca2+-free cell culture medium (J-MEM) containing 1% casein, prior to homogenization in a TEDG buffer (50 mm Tris-maleate buffer, pH 7.4; 1.5 mm EDTA; 2.0 mm DTT; and 10% (vv) glycerol) containing 0.5 mm phenylmercuric sulfonyl fluoride (PMSF) and 1.0 mm sodium azide. Na2MoO4 (final concentration, 20 mm) was added to homogenate fractions which then were incubated at 0–4 °C for 1–5 days with a saturating concentration of 3H-R-1881 (20 nm) in the absence and presence of 2 μm radioinert R-1881. Heparin (200 μg/ml) was added to the incubation medium to “solubilize” nuclear chromatin. Free and bound R-1881 were separated by a hydroxylapatite (HAP) batch procedure. Using these conditions, it has been found that (i) incubation periods of 72–96 h at 0–4 °C were required to achieve maximal specific exchange binding of 3H-R-1881 (Bmax) in total homogenates from normal intact rats. Heparin addition (200 μg/ml) did not change Bmax and had little or no effect on the rate of exchange. Mean Bmax was 6.7 ± 1.6 (SD) pmol 3H-R-1881/mg DNA. R-1881 exchange at 24 h of incubation was only about 40% of Bmax. Nonspecific binding, a small fraction (<10%) of Bmax, was near maximal at 2 h. Incubation at 15 °C gave similar R-1881 exchange values to those obtained at 0 °C during the first 24 h, but at 48 h and thereafter R-1881 binding markedly decreased. These Bmax values in total homogenates of normal intact RVP are about 2.5 times greater than the AR values obtained in 1-day castrated rats, when compared on a DNA basis, (ii) Addition of gelatin (0.25%) to inhibit seminin activity had no effect on Bmax. Deletion of either MoO42? or PMSF from the standard exchange medium reduced Bmax values ~20%; if both PMSF and MoO42? were deleted, Bmax was reduced to a greater extent (~35%). Bmax was reduced (40%) when homogenates were prepared without preliminary J-MEM:casein pretreatment and incubated in standard exchange medium with PMSF and MoO42?. (iii) Despite AR stabilization by MoO42? and inhibition of protease activities during exchange incubation, AR inactivation increased exponentially, so that the maximal 3H-R-1881 binding value achieved at 96 h was estimated to represent about 50% of the AR sites originally present. (iv) The binding sites in total homogenates occupied by 3H-R-1881 at 24, 72, and 96 h of exchange exhibited steroid specificity characteristics of AR, as determined by competition studies with a wide variety of steroid hormones and analogues. Scatchard plots of 3H-R-1881 exchange binding in total homogenates of normal intact RVP incubated for 72 or 96 h indicated a single class of affinity sites with apparent Kd of 5 to 6 nm. (v) Sucrose density gradient centrifugation of homogenates incubated for 72 or 96 h showed that the specific 3H-R-1881 binding sites were distributed in two broad peaks associated with low-molecular-weight components. One with S value ~3.5 may be “activated” AR; the other near the top of the gradient (S < 1.6) may include meroreceptor forms of AR.  相似文献   

20.
Stoichiometry of sulphide and intracellular sulphur oxidation in connection with CO2 fixation was studied inChromatium okenii. The equipment used was a special stirred cuvette with a rapid-sampling arrangement, which allowed short-time experiments with illuminated bacterial suspensions under anaerobic conditions. Turnover of the sulphur compounds is controlled by a linear CO2 fixation rate which amounts to 0.069µmoles of CO2/min mg of cell protein at light saturation. Van Niel's equations for bacterial photosynthesis could be confirmed for short periods under the condition that sulphate is produced during increase of intracellular sulphur; i.e., oxidation of sulphide and of intracellular sulphur do not occur consecutively but simultaneously. The full oxidation rate of intracellular sulphur starts after complete consumption of sulphide. The time during which sulphide is oxidized to intracellular sulphur amounts to 1/3–1/4 of the time necessary for the complete quantitative oxidation of the sulphide to sulphate.  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