首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Phosvitin, a phosphoprotein known as an iron-carrier in egg yolk, binds almost all the yolk iron. In this study, we investigated the effect of phosvitin on Fe(II)-catalyzed hydroxyl radical (?OH) formation from H2O2 in the Fenton reaction system. Using electron spin resonance (ESR) with 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) and deoxyribose degradation assays, we observed by both assays that phosvitin more effectively inhibited ?OH formation than iron-binding proteins such as ferritin and transferrin. The effectiveness of phosvitin was related to the iron concentration, indicating that phosvitin acts as an antioxidant by chelating iron ions. Phosvitin accelerates Fe(II) autoxidation and thus decreases the availability of Fe(II) for participation in the ?OH-generating Fenton reaction. Furthermore, using the plasmid DNA strand breakage assay, phosvitin protected DNA against oxidative damage induced by Fe(II) and H2O2. These results provide insight into the mechanism of protection of the developing embryo against iron-dependent oxidative damage in ovo.  相似文献   

2.
Human serum heme–albumin (HSA–heme–Fe) displays reactivity and spectroscopic properties similar to those of heme proteins. Here, the nitrite reductase activity of ferrous HSA–heme–Fe [HSA–heme–Fe(II)] is reported. The value of the second-order rate constant for the reduction of $ {\text{NO}}_{2}^{ - } $ to NO and the concomitant formation of nitrosylated HSA–heme–Fe(II) (i.e., k on) is 1.3 M?1 s?1 at pH 7.4 and 20 °C. Values of k on increase by about one order of magnitude for each pH unit decrease between pH 6.5 to 8.2, indicating that the reaction requires one proton. Warfarin inhibits the HSA–heme–Fe(II) reductase activity, highlighting the allosteric linkage between the heme binding site [also named the fatty acid (FA) binding site 1; FA1] and the drug-binding cleft FA2. The dissociation equilibrium constant for warfarin binding to HSA–heme–Fe(II) is (3.1 ± 0.4) × 10?4 M at pH 7.4 and 20 °C. These results: (1) represent the first evidence for the $ {\text{NO}}_{2}^{ - } $ reductase activity of HSA–heme–Fe(II), (2) highlight the role of drugs (e.g., warfarin) in modulating HSA(–heme–Fe) functions, and (3) strongly support the view that HSA acts not only as a heme carrier but also displays transient heme-based reactivity.  相似文献   

3.
We have developed a rapid, inexpensive, and reliable assay to determine ascorbate using potassium ferricyanide as spectroscopic probe reagent. In this assay, Fe(III) was deoxidized to Fe(II) by ascorbate at pH 4.0 and then Fe(II) reacted with potassium ferricyanide to form a blue product, soluble Prussian blue (KFeIII[FeII(CN)6]). The absorbance of this product was monitored over time using a spectrophotometer at an absorption maximum of 735 nm and the amount of ascorbate can be calculated based on absorbance. A good linear relationship of the concentration of ascorbate versus absorbance was observed, and the linear regression equation was A = −0.01911 + 0.16208C (μg/ml). Moreover, the apparent molar absorption coefficient of indirect determination of ascorbate was 2.85 × 104 L/mol·cm. To demonstrate the usefulness of this assay, it was used to determine ascorbate in different samples, and we particularly investigated the uptake of ascorbate and ascorbate phosphate in osteoblasts. We found similar plateau levels of intracellular ascorbate at 24 h for ascorbate and ascorbate phosphate. The assay was robust for a variety of samples, including orange juice, fruits, and swine plasma. The assay was quick and very economical and provides results with uncertainties on the order of only 5%.  相似文献   

4.
Using [3H]naloxone at a concentration of 4.5 nm , the potent opiate agonist etorphine as well as the potent antagonist diprenorphine displace only about 75% of specific naloxone binding P2 fractions from rat whole forebrain, without additive effect. Several other opiates and antagonists completely displace specific naloxone binding. This indicates that etorphine and diprenorphine specifically bind to one and the same naloxone binding site (type I) while leaving another naloxone binding site (type II) unaffected. Type I binding sites are much more thermo-labile than type II. [3H]Naloxone binding to type I sites is unaffected by incubation temperature in the range 10 to 25°C. while binding type II sites decreases rapidly with increasing incubation temperature, no specific type II binding being detectable at or above 20°C. The two naloxone receptor types also differ with respect to pH dependence, and affinity for naloxone with types I and II having affinity constants (Kd) of 2 and 16 nm , respectively, at 0°C. The two binding sites have different regional distributions with high relative levels of type II receptors in cerebellum and low relative levels in pons-medulla and striatum. In whole rat brain there are about 4 times as many type II receptors as type I. These results suggest that naloxone and several other opiate agonists and antagonists bind to two distinct receptor types which are probably not agonist/antagonist aspects of the same receptor.  相似文献   

5.
The influence of FeEDDHA (0, 0.2 and 2 μg Fe g−1 soil) and NaH2PO4·H2O (0 and 120 μg Pg−1 soil) on the growth of two Fe-ineffective soybean (Glycine max L. Merr.) varieties (anoka and T203) on a calcareous soil at two soil temperatures (16 and 24°C) was compared under greenhouse conditions. The two soybean varieties differed in the following respects: (a) T203 accumulated smaller concentrations of Fe in washed tops than Anoka under comparable conditions; (b) T203 was more susceptible to Fe deficiency and its accentuation by high levels of fertilizer P than Anoka; (c) T203 accumulated lower quantities of Mn in tops than Anoka under comparable conditions; (d) T203, but not Anoka, developed Mn deficiency symptoms when treated with P and 2 μg Fe g−1 at 16°C. Fe deficiency was more severe in both varieties at the higher soil temperature due apparently to: (a) greater plant concentration of P in tops at 24°C; and/or (b) an increased rate of plant growth and greater dilution of Fe in young tissue at 24°C. Foliar P concentration was increased much more than foliar Fe concentration by an increase in soil temperature. Severely Fe deficient T203 plants grown without FeEDDHA at 24°C accumulated less foliar Mn than their FeEDDHA counterparts. Comparisons of Fe effectiveness of various soybean cultivars based on relative responses to FeEDDHA can be influenced by differential effects on Mn nutrition.  相似文献   

6.
Thermosipho globiformans (rod-shaped thermophilic fermenter) and Methanocaldococcus jannaschii (coccal hyperthermophilic hydrogenotrophic methanogen) established H2-mediated syntrophy at 68 °C, forming exopolysaccharide-based aggregates. Electron microscopy showed that the syntrophic partners connected to each other directly or via intercellular bridges made from flagella, which facilitated transfer of H2. Elemental sulfur (S0) interrupted syntrophy; polysulfides abiotically formed from S0 intercepted electrons that were otherwise transferred to H+ to produce H2, resulting in the generation of sulfide (sulfur respiration). However, Fe(III) oxides significantly reduced the interruption by S0, accompanied by stiffening of Fe(II) sulfides produced by the reduction of Fe(III) oxides with the sulfur respiration-generated sulfide. Sea sand replacing Fe(III) oxides failed to generate stiffening or protect the syntrophy. Several experimental results indicated that the stiffening of Fe(II) sulfides shielded the liquid from S0, resulting in methane production in the liquid. Field-emission scanning electron microscopy showed that the stiffened Fe(II) sulfides formed a network of spiny structures in which the microorganisms were buried. The individual fermenter rods likely produced Fe(II) sulfides on their surface and became local centers of a core of spiny structures, and the connection of these cores formed the network, which was macroscopically recognized as stiffening.  相似文献   

7.
The pH of the solution along with chelation and consequently coordination of iron regulate its reactivity. In this study we confirmed that, in general, the rate of Fe(II) autoxidation increases as the pH of the solution is increased, but chelators that provide oxygen ligands for the iron can override the affect of pH. Additionally, the stoichiometry of the Fe(II) autoxidation reaction varied from 2:1 to 4:1, dependent upon the rate of Fe(II) autoxidation, which is dependent upon the chelator. No partially reduced oxygen species were detected during the autoxidation of Fe(II) by ESR using DMPO as the spin trap. However, upon the addition of ethanol to the assay, the DMPO:hydroxyethyl radical adduct was detected. Additionally, the hydroxylation of terephthalic acid by various iron-chelator complexes during the autoxidation of Fe(II) was assessed by fluorometric techniques. The oxidant formed during the autoxidation of EDTA:Fe(II) was shown to have different reactivity than the hydroxyl radical, suggesting that some type of hypervalent iron complex was formed. Ferrous iron was shown to be able to directly reduce some quinones without the reduction of oxygen. In conclusion, this study demonstrates the complexity of iron chemistry, especially the chelation of iron and its subsequent reactivity.  相似文献   

8.
A simple and rapid method is described for determining Pi by spectrophotometric measurement of a soluble complex of phosphomolybdic acid and Cirrasol ALN-WF, a non-ionic detergent formerly known as Lubrol W. The measured complex has a molar extinction coefficient of 4.59 · 103 at 390 nm and little interference is found with relatively high concentrations of chelating agents, salts, and other compounds which interfere with most other Pi assays. Linearity is observed in the range 0–1.2 μmoles Pi and developed assay samples are stable for 8 h at 20 °C or 24 h at 4 °C. The method is suitable for use in the presence of moderate concentrations of protein or ATP.After suitable modification the assay can be used at pH 4.0. Sensitivity is reduced at this pH (εM, 390nm = 2.79 · 103) but linearity is maintained up to 1 μmole Pi and the coloured complex is stable for 4 h at 20 °C. The pH-4 procedure is suitable for measurement of Pi in the presence of very labile phosphate esters such as creatine phosphate.The phosphomolybdic acid-Cirrasol complex can be reduced at ambient temperature in both the above systems. A blue complex results with εM, 820nm of 9.9 · 103 at pH 4.0, and 1.8 · 104 under more acidic conditions.  相似文献   

9.
There are increasing numbers of exceptions to a central tenet in cryobiology that low-molecular-weight protective solutes such as glycerol must permeate cells in high concentration in order to protect them from freezing injury. To test this supposition, it is necessary to determine the amount of solute that has permeated a cell prior to freezing. The amount in human red cells was estimated from the flux equation dsdt = PγA[(activity external solute) — (activity internal solute)]. Solving the equation required knowledge of Pγ the permeability constant for the solute. Estimates of Pγ for glycerol were made in two ways: (i) by measuring the time to 50% hemolysis of human red cells suspended in 1 or 2 m solutions of glycerol that were hypotonic with respect to NaCl, and (ii) by measuring the time required for red cells in 1 or 2 m solutions of glycerol in isotonic saline-buffer to undergo osmotic shock upon tenfold dilution with isotonic saline-buffer. The measurements were made at 0 and 20 °C. The values of Pγ were about 2.5 × 10?4 cm/min at 20 °C and about 0.9 × 10?4 cm/min at 0 °C. The difference corresponds to an activation energy of 7.2 kcal/mole. These values of Pγ are 100 to 600 times higher than those for glycerol permeation in the bovine erythrocyte. The values of P were relatively unaffected by whether calculations were based on classical or irreversible thermodynamics and by the choice of concentration units in the flux equations. Calculations of the kinetics of glycerol entry using these P values showed that the concentration of intracellular glycerol reaches 90% of equilibrium in 1.2 min at 0 °C and in 0.6 min at 20 °C. The osmolal ratio of intracellular glycerol to intracellular nonpermeating solutes reaches 90% of equilibrium in 7 min at 0 °C and in 3.2 min at 20 °C.  相似文献   

10.
Quinolinate (pyridine-2,3-dicarboxylic acid, Quin) is a neurotoxic tryptophan metabolite produced mainly by immune-activated macrophages. It is implicated in the pathogenesis of several brain disorders including HIV-associated dementia. Previous evidence suggests that Quin may exert its neurotoxic effects not only as an agonist on the NMDA subtype of glutamate receptor, but also by a receptor-independent mechanism. In this study we address ability of ferrous quinolinate chelates to generate reactive oxygen species. Autoxidation of Quin-Fe(II) complexes, followed in Hepes buffer at pH 7.4 using ferrozine as the Fe(II) detector, was found to be markedly slower in comparison with iron unchelated or complexed to citrate or ADP. The rate of Quin-Fe(II) autoxidation depends on pH (squared hydroxide anion concentration), is catalyzed by inorganic phosphate, and in both Hepes and phosphate buffers inversely depends on Quin concentration. These observations can be explained in terms of anion catalysis of hexaaquairon(II) autoxidation, acting mainly on the unchelated or partially chelated pool of iron. In order to follow hydroxyl radical generation in the Fenton chemistry, electron paramagnetic resonance (EPR) spin trapping with 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) was employed. In the mixture consisting of 100 mM DMPO, 0.1 mM Fe(II), and 8.8 mM hydrogen peroxide in phosphate buffer pH 7.4, 0.5 mM Quin approximately doubled the yield of DMPO-OH adduct, and higher Quin concentration increased the spin adduct signal even more. When DMPO-OH was pre-formed using Ti3+/hydrogen peroxide followed by peroxide removal with catalase, only addition of Quin-Fe(II), but not Fe(II), Fe(III), or Quin-Fe(III), significantly promoted decomposition of pre-formed DMPO-OH. Furthermore, reaction of Quin-Fe(II) with hydrogen peroxide leads to initial iron oxidation followed by appearance of iron redox cycling, detected as slow accumulation of ferrous ferrozine complex. This phenomenon cannot be abolished by subsequent addition of catalase. Thus, we propose that redox cycling of iron by a Quin derivative, formed by initial attack of hydroxyl radicals on Quin, rather than effects of iron complexes on DMPO-OH stability or redox cycling by hydrogen peroxide, is responsible for enhanced DMPO-OH signal in the presence of Quin. The present observations suggest that Quin-Fe(II) complexes display significant pro-oxidant characteristics that could have implications for Quin neurotoxicity.  相似文献   

11.
3-Hydroxykynurenine (3-OHKyn) is a tryptophan metabolite that is readily autoxidised to products that may be involved in protein modification and cytotoxicity. The oxidation of 3-OHKyn has been studied here with a view to characterising the major products as well as determining their relative rates of formation and the role that H2O2 and hydroxyl radical (HO·) may play in modifying the autoxidation process. Oxidation of 3-OHKyn generated several compounds. Xanthommatin (Xan), formed by the oxidative dimerisation of 3-OHKyn, was the major product formed initially. It was, however, found to be unstable, particularly in the presence of H2O2, and degraded to other products including the p-quinone, 4,6-dihydroxyquinolinequinonecarboxylic acid (DHQCA). A compound that has a structure consistent with that of hydroxy-xanthommatin (OHXan) was also formed in addition to at least two minor species that we were unable to identify. Hydrogen peroxide was formed rapidly upon oxidation of 3-OHKyn, and significantly influenced the relative abundance of the different autoxidation species. Increasing either pH (from pH 6 to 8) or temperature (from 25°C to 35°C) accelerated the rate of autoxidation but had little impact on the relative abundance of the autoxidation species. Using electron paramagnetic resonance (EPR) spectroscopy, a clear phenoxyl radical signal was observed during 3-OHKyn autoxidation and this was attributed to xanthommatin radical (Xan·). Hydroxyl radicals were also produced during 3-OHKyn autoxidation. The HO· EPR signal disappeared and the Xan· EPR signal increased when catalase was added to the autoxidation mixture. The HO· did not appear to play a role in the formation of the autoxidation products as evidenced using HO· traps/scavengers. We propose that the cytotoxicity of 3-OHKyn may be explained by both the generation of H2O2 and by the formation of reactive 3-OHKyn autoxidation products such as the Xan· and DHQCA.  相似文献   

12.
Lactoperoxidase (LPO) is a hemeprotein catalyzing the oxidation of thiocyanate and I? into antimicrobials and small aromatic organics after being itself oxidized by H2O2. LPO is excreted by the lungs, mammary glands, found in saliva and tears and protects mammals against bacterial, fungal and viral invasion. The Fe(II) form binds CO which inactivates LPO like many other hemeproteins. We present the 3-dimensional structure of CO?CLPO at 2.0? resolution and infrared (IR) spectra of the iron-bound CO stretch from pH?3 to 8.8?at 1 cm?1 resolution. The observed Fe?CC?CO bond angle of 132° is more acute than the electronically related Fe(III), CN?CLPO with a Fe?CC?CN angle of 161°. The orientations of the two ligands are different with the oxygen of CO pointing towards the imidazole of distal His109 while the nitrogen of CN points away, the Fe(II) moves towards His109 while the Fe(III) moves away; both movements are consistent with a hydrogen bond between the distal His109 and CO, but not to the nitrogen of CN?CLPO. The IR spectra of CO?CLPO exhibit two major CO absorbances with pH dependent relative intensities. Both crystallographic and IR data suggest proton donation to the CO oxygen by His109 with a pK ?? 4; close to the pH of greatest enzyme turnover. The IR absorbance maxima are consistent with a first order correlation between frequency and Fe(III)/Fe(II) reduction potential at pH?7; both band widths at half-height correlate with electron density donation from Fe(II) to CO as gauged by the reduction potential.  相似文献   

13.
In comparison with myoglobin molecule as a reference, we have studied the autoxidation rate of human oxyhemoglobin (HbO2) as a function of its concentration in 0.1 M buffer at 35°C and in the presence of 1 mM EDTA. At pH 6.5, HbA showed a biphasic autoxidation reaction that can be described completely by a first-order rate equation containing two rate constants — kf, for fast autoxidation of the α-chain, and ks, for slow autoxidation of the β-chain, respectively. When tetrameric HbO2 was dissociated into αβ-dimers by dilution, the value of kf increased markedly to an extent comparable with the autoxidation rate of horse heart oxymyoglobin (MbO2). The rate constant ks, on the other hand, was found to remain at an almost constant value over the whole concentration range from 1.0 × 10−3 M to 3.2 × 10−6 M in heme. At pH 8.5 and pH 10.0, however, the autoxidation of HbO2 was monophasic, and no enhancement in the rate was observed by diluting hemoglobin solutions. Taking into consideration the effects of 2,3-diphosphoglyceric acid and chloride anion on the autoxidation rate of HbO2, we have characterized the differential susceptibility of the α- and β-chains to the autoxidation reaction in aqueous solution.  相似文献   

14.

Background

The 2-deoxyribose (2-DR) degradation assay is a widely used test for determining anti/pro-oxidant properties of molecules and plant extracts. Most reports use reaction blanks omitting 2-DR or thiobarbituric acid (TBA). However, when studying Fe(II)-mediated reactions, we verified that these blanks are not appropriate. Fe(III) – a product of these reactions – causes a relevant artifact in the assay, where 2-DR is oxidized by Fe(III).

Method

2-DR degradation was determined at 532 nm as TBA-reactive substances.

Results and conclusion

HPLC determinations indicated that Fe(III) added after or before TBA generates considerable amounts of malondialdehyde (2-DR degradation product) in comparison with assays employing Fenton reagents or Fe(II) autoxidation. Addition of catalase and thiourea has no effect on Fe(III)-induced 2-DR degradation indicating lack of ROS involvement. This Fe(III)-mediated 2-DR damage is dependent on iron and 2-DR concentrations, but not on H2O2, buffer composition or iron-chelators. Depending on the assay conditions Fe(III)-interference accounts for 20% to 90% of 2-DR degradation mediated by Fe(II).

Significance

A new reaction blank is proposed herein–based on the use of Fe(III)–for the assay. The lack of such correction has caused the underestimation of antioxidant capacity of various compounds in many studies in the last 2 decades.  相似文献   

15.
A modification of a colorimetric assay (1) discriminating galactosamine and glucosamine in mixtures is presented. The procedure employs acetylation of the amino sugars at 25°C instead of at 0–1°C, resulting in five times the color intensity for galactosamine with no interference due to glucosamine. The assay of HCl hydrolysates of glycosaminoglycans and glycoproteins is possible in solutions that are less than 0.1 n.  相似文献   

16.
Abstract

Chemical and enzymatic footprinting experiments have made it possible to identify protein binding sites in DNA and RNA and to localize structural differences within nucleic acids to a resolution of a single base pair. We show here that by combining three reagents, Fe(II)·EDTA2·Fe(II)·EDDA and Fe2+, differential maps of sites in RNA that vary in their local conformation and/or charge can be constructed. Comparison of profiles with respect to controls in the absence of a counterion such as Mg2+ allows analysis of sites responsive to tertiary structure. A single site that is labile to metals such as Pb2+ exists in tRNAPhe and a number of other tRNA's; this site is hyper-reactive to Fe(II), but not to the other probes. Scission induced by the neutral complex, Fe(II)·EDDA, offers the most general measure of surface accessibility, since its distribution about the target molecule is insensitive to charge. Enhanced cleavage by Fe(II) relative to the other agents is detected at several adjacent sites in 5S RNA consistent with conformational mobility. Protection at a series of positions in the arm formed by loops E and D with helix IV suggests further that at low temperature this arm interacts with loop A and helix I.  相似文献   

17.
Summary A previous study described a cytoplasmic, transferrin (Tf)-free, iron (Fe) pool that was detected only when cells were mechanically detached from the culture substratum at 4°C, after initial incubation with59Fe-125I-Tf at 37°C (Richardson and Baker, 1992a). The release of this internalized59Fe could be markedly reduced if the cells were treated with proteases or incubated at 37°C prior to detachment. The present study was designed to characterize this Fe pool and understand the mechanism of its release. The results show that cellular59Fe release increased linearly as a function of preincubation time with59Fe-Tf subsequent to mechanical detachment at 4°C using a spatula. These data suggest that the59Fe released was largely composed of end product(s) and was not an “intermediate Fe pool.” When the Fe(II) chelator, dipyridyl (DP), was incubated with59Fe-Tf and the cells, it prevented the accumulation of59Fe that was released following mechanical detachment at 4°C. Other chelators had much less effect on the proportion of59Fe released. Examination of the59Fe released showed that after a 4-h preincubation with59Fe-Tf, approximately 50% of the59Fe was present in ferritin. These data indicate that mechanical detachment of cells at 4°C resulted in membrane disruptions that allow the release of high M, molecules. Moreover, electron microscopy studies showed that detachment of cells from the substratum at 4°C resulted in pronounced membrane damage. In contrast, when cells were detached at 37°C, or at 4°C after treatment with pronase, membrane damage was minimal or not apparent. These results may imply that temperature-dependent processes prevent the release of intracellular contents on membrane wounding, or alternatively, prevent wounding at 37°C. The evidence also indicates that caution is required when interpreting data from expriments where cells have been mechanically detached at 4°C.  相似文献   

18.
Neutrophilic Fe(II) oxidizing microorganisms are found in many natural environments. It has been hypothesized that, at low oxygen concentrations, microbial iron oxidation is favored over abiotic oxidation. Here, we compare the kinetics of abiotic Fe(II) oxidation to oxidation in the presence of the bacterium Leptothrix cholodnii Appels isolated from a wetland sediment. Rates of Fe(II) oxidation were determined in batch experiments at 20°C, pH 7 and oxygen concentrations between 3 and 120 μmol/l. The reaction progress in experiments with and without cells exhibited two distinct phases. During the initial phase, the oxygen dependency of microbial Fe(II) oxidation followed a Michaelis-Menten rate expression (KM = 24.5 ± 10 μmol O2/l, vmax = 1.8 ± 0.2 μmol Fe(II)/(l min) for 108 cells/ml). In contrast, abiotic rates increased linearly with increasing oxygen concentrations. At similar oxygen concentrations, initial Fe(II) oxidation rates were faster in the experiments with bacteria. During the second phase, the accumulated iron oxides catalyzed further oxidative iron precipitation in both abiotic and microbial reaction systems. That is, abiotic oxidation also dominated the reaction progress in the presence of bacteria. In fact, in some experiments with bacteria, iron oxidation during the second phase proceeded slower than in the absence of bacteria, possibly due to an inhibitory effect of extracellular polymeric substances on the growth of Fe(III) oxides. Thus, our results suggest that the competitive advantage of microbial iron oxidation in low oxygen environments may be limited by the autocatalytic nature of abiotic Fe(III) oxide precipitation, unless the accumulation of Fe(III) oxides is prevented, for example, through a close coupling of Fe(II) oxidation and Fe(III) reduction.  相似文献   

19.
A detailed picture of the orientation and restricted motion of the cholestane spin label (3-spiro-doxyl-5α-cholestane) in planar multibilayers of dipalmitoylphosphatidylcholine and dimyristoylphosphatidylcholine has been recorded by simultaneous simulation of ESR spectra obtained with the magnetic field parallel and perpendicular to the bilayers (Shimoyama, Y., Eriksson, L.E.G. and Ehrenberg, A. (1978) Biochim. Biophys. Acta 508, 213–235). The analysis has been made over the temperature range ?30°C to 60°C on samples containing 20 to 22% water. At low temperatures the cholestane spin label is tilted with respect to the lipid bilayer normal by an angle of approx. 30° which disappears at the pretransition. In this low temperature range the restricted twisting motion has an activation energy of 5.5 kJ·mol?1. Above the main transition the twisting motion is unrestricted and has the activation energy 20 kJ·mol?1. From below the pretransition to above the main transition the velocity of the twisting motion increases by an order of magnitude. The amplitude of the wobbling motion increases abruptly from 0° to 35° at the main transition.  相似文献   

20.
The present study examines the kinetics and mechanism of the system [FePDTA(OH)]2− + 5CN ⇌ [Fe(CN)5OH]3− + PDTA4− at pH= 11.0±0.02, I= 0.25 M and temperature = 25 ± 0.1 °C. The reaction has been studied spectrophotometrically at 395 nm (λmax of [Fe(CN)5OH]3−). The data show that the reaction has three distinguishable stages; the first stage is formation of [Fe(CN)5OH]3−, the second is conversion of [Fe(CN)5OH)]3− to [Fe(CN)6]3− and last is reduction of [Fe(CN)6]3− to [Fe(CN)6]3− by the released ligand, viz., PDTA. The first reaction shows variable order dependence on cyanide concentration, one at high cyanide concentration and two at low cyanide concentration. The second reaction exhibits first order dependence on the concentration of [Fe(CN)5OH]3− as well as cyanide. The reverse reaction between [Fe(CN)5OH]3− and PDTA is first order in [Fe(CN)5OH]3− and PDTA, and inverse first order in cyanide. On the basis of forward and reverse rate studies, a five-step mechanism has been proposed for the first reaction.  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