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.
The reaction of HbO2 with phenols to produce metHb shows inverse rate dependence upon [H+], direct dependence upon [HbO2] and [phenol], and a rate that correlates with the electron donor characteristics of the reagents. Thus, the availability of an electron from an external agent permits facile reduction of O2 to O2= and the reaction of HbO2 with phenols gives rise to metHb and peroxide as reaction products. In contrast, with nucleophiles such as azide O2 is displaced as superoxide. Since reduction of bound O2 is seen to occur only by reductive displacement or by reaction with a single electron donor, Hb apparently owes its normal resistance to autoxidation to the isolation of the binding site from electron donors and nucleophiles and not to an unique kind of iron-O2 bonding. Such reasoning explains the effects of structural abnormality that render M-type Hbs susceptible to oxidation. Also the oxidation of HbO2 upon exposure to “oxidant drugs” is explicable in terms of the drugs acting as one electron reducing agents towards bound dioxygen.  相似文献   

2.
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   

3.
HbA O2 reacts readily with FeII(CN)5H2O3? to form aquometHb and peroxide via a second order process: rate=k[HbO2][FeII(CN)5H2O3?]. A slight enchancement in the rate of metHb formation due to the H2O2 produced can be prevented by addition of catalase. The reaction is free from complications exhibited by other reductants. The hexacyanide, ferrocyanide, reacts with HbA O2 but at only ca. 0.02% the rate and with formation of cyanometHb. Reductants such as phenols and sulfa drugs may produce radicals that can enter into side reactions. FeII(CN)5H2O3? shows promise as an effective probing reagent for the characterization of H2O2 production from oxygenated heme and other proteins.  相似文献   

4.
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   

5.
Various hemoglobin derivatives have been labeled at the Cys-β93 residue with a bulky and “strongly immobilized” nitroxide maleimide (I) and a smaller, more flexible and “weakly immobilized” nitroxide iodoacetamide (II) and crystallized. The angular dependence of the paramagnetic resonance of the spin-label was measured for the ab, ac1 and bc1 planes at 298 K and 77 K for spin-labeled crystals of Oxyhemoglobin, methemoglobin fluoride, and methemoglobin azide. In the case of the methemoglobin crystals, the angular variation of the heme resonance was also monitored at 77 K. From the hyperfine splitting data, the spin-label I was found to assume specific orientations at both temperatures with some motional narrowing at 298 K. Spin-label II is specifically oriented only at room temperature but is frozen at 77 K in random orientations. Oxyhemoglobin labeled with I (I-HbO22) has the most prominent spin-label orientation (zb, xa) and the less abundant spin-labels with (zb ± 15 °) (Ohnishi et al., 1966). The corresponding spin-label orientations for I-Hb+ F? are (z∥a, x∥c1) and (z∥c1, x∥a). Crystals of I-Hb+ N?3 have spin-labels oriented along angular directions similar, but not identical to those of I-Hb+F?. Therefore, there are probably significant peptide segmental displacements when HbO2 is oxidized to methemoglobins. At 25 °C II-Hb+ N?3 has spin-label orientations not too different from those in I-Hb+ N?3, whereas in HbO2 the two spin-labels show significant differences in their orientations.  相似文献   

6.
The nonenzymatic isoprostane pathway of lipid peroxidation of polyunsaturated fatty acids results in formation of products, termed isoprostanes, which have very large positional and stereo isomerism, possess various biological activities, produce adducts with proteins, and thus contribute to pathogeneses of the agedependent diseases. However, it was unclear what mechanism drives this type of lipid autoxidation, and why the products have very large isomerism. We propose a mechanism when perhydroxyl radicals (HO2?) react with polyunsaturated fatty acids in the hydrophobic milieu of membranes. In the membrane HO2? initiates a chain of reactions with formation first H2O2, which undergoes homolytic fission producing two ?OH radicals, thus very rapidly abstracting three H atoms from a polyunsaturated fatty acid. As a result, the HO2? molecule is converted to two molecules of water, and the molecule of a polyunsaturated fatty acid loses two double bonds, becomes highly unstable and undergoes peroxidation and random intramolecular re-arrangements causing a very large isomerism of the final products. The extremely high reactivity of ?2 with polyunsaturated fatty acids is the cause of very subtle and slow accumulation of damages in the membrane and membrane associated proteins, even though the concentration of ?2 relative to superoxide radical may be very low.  相似文献   

7.
The effect of different co-anions on the formation and aggregation of the ordered structure of the anionic polysaccharide kappa-carrageenan has been investigated by optical rotation, differential scanning calorimetry, and halide-n.m.r. spectroscopy. The mid-point temperature (Tm) of the disorder—order transition increases systematically with the Hofmeister number for the anion through the lyotropic series SO42? < F? < Cl? < Br? < NO3? < I? < SCN? when salt concentration and cation (Me4N+ or K+) are held constant. A corresponding increase is observed in transition enthalpy (ΔHcal) and entropy (ΔScal). Helix—helix aggregation (as indicated by turbidity, gel formation, and hysteresis between heating and cooling scans) also shows a systematic dependence on the Hofmeister number for the anion, but in the opposite sense. Thus, with tetramethylammonium as the sole counterion present, clear solutions with no thermal hysteresis in the order—disorder transition are observed at all temperatures with I?, Br?, NO3?, or Cl? as co-anion, whereas weak, turbid gels with significant thermal hysteresis between melting and setting are formed in the presence of SO42?, and to a lesser degree F?. With K+ as counterion, a similar regular progression is observed through the anion lyotropic series from rapid formation of very turbid gels in the presence of F?, to very slow development of clear gels with I? or SCN?. In agreement with previous studies, an increase in 127I-n.m.r. linewidth was observed on conformational ordering of kappa-carrageenan (Me4N+ salt form) in the presence of Me4NI. However, closely similar behaviour was observed for 35Cl and 81Br, indicating a simple charge-cloud interaction rather than the specific site-binding of I? which has previously been suggested.  相似文献   

8.
Wastewater treatment plants are known to be important point sources for nitrous oxide (N2O) in the anthropogenic N cycle. Biofilm based treatment systems have gained increasing popularity in the treatment of wastewater, but the mechanisms and controls of N2O formation are not fully understood. Here, we review functional groups of microorganism involved in nitrogen (N) transformations during wastewater treatment, with emphasis on potential mechanism of N2O production in biofilms. Biofilms used in wastewater treatment typically harbour aerobic and anaerobic zones, mediating close interactions between different groups of N transforming organisms. Current models of mass transfer and biomass interactions in biofilms are discussed to illustrate the complex regulation of N2O production. Ammonia oxidizing bacteria (AOB) are the prime source for N2O in aerobic zones, while heterotrophic denitrifiers dominate N2O production in anoxic zones. Nitrosative stress ensuing from accumulation of NO2 ? during partial nitrification or denitrification seems to be one of the most critical factors for enhanced N2O formation. In AOB, N2O production is coupled to nitrifier denitrification triggered by nitrosative stress, low O2 tension or low pH. Chemical N2O production from AOB intermediates (NH2OH, HNO, NO) released during high NH3 turnover seems to be limited to surface-near AOB clusters, since diffusive mass transport resistance for O2 slows down NH3 oxidation rates in deeper biofilm layers. The proportion of N2O among gaseous intermediates (NO, N2O, N2) in heterotrophic denitrification increases when NO or nitrous acid (HNO2) accumulates because of increasing NO2 ?, or when transient oxygen intrusion impairs complete denitrification. Limited electron donor availability due to mass transport limitation of organic substrates into anoxic biofilm zones is another important factor supporting high N2O/N2 ratios in heterotrophic denitrifiers. Biofilms accommodating Anammox bacteria release less N2O, because Anammox bacteria have no known N2O producing metabolism and reduce NO2 ? to N2, thereby lowering nitrosative stress to AOB and heterotrophs.  相似文献   

9.
Enhanced electrochemical resolution of anodic processes is possible in the presence of [N(nBu)4][B(C6F5)4], 1, as supporting electrolyte over that obtained in the presence of [N(nBu)4][PF6]. By changing the anion of the supporting electrolyte to a salt having [B(C6F5)4], anions, electrochemical processes of especially cationic analytes can benefit. Thus, the redox chemistry of 0.5 mmol dm−3 solutions of [Ru2(μ-FcCOO)4·(CH3CH2OH)2][PF6], 2, Fc = ferrocenyl, in CH2Cl2/[N(nBu)4][B(C6F5)4] were found to involve four well-resolved ferrocenyl-based electrochemical reversible redox processes as well as reduction of RuIII-RuII. At 1.0 mmol dm−3 concentrations of 2, or in the presence of [N(nBu)4][PF6], the four ferrocenyl processes coalesced into only two waves as a result of (Fc+)?() ion paring. Seventeen of the possible 18 one-electron transfer processes of the biscadmium trisphthalocyaninato complex [Cd2{Pc(C6H13)8}3], 3, could be observed in THF/[N(nBu)4][B(C6F5)4], but the electrochemical window of CH2Cl2/[N(nBu)4][B(C6F5)4] only allowed detection of 15 of these processes. Although reduction processes were unaffected, THF solvation leading to species such as (3n+)(THF)x with 1 ? n ? 4 and x ? 1 as well as ion pair formation of the type (3n+)?() prevented good resolution of oxidation processes. The CH2Cl2/[N(nBu)4][B(C6F5)4] system also allowed detection of reversible one-electron transfer ferrocenyl (Fc/Fc+) and ruthenocenyl-based (Rc/Rc+) processes for both enol and keto isomers of the β-diketone FcCOCH2CORc, 4, Rc = ruthenocenyl. In CH3CN/[N(nBu)4][PF6], the ruthenocenyl moiety was oxidised to a RuIV species.  相似文献   

10.
Two new zinc phosphates, [C8N4H26][Zn2(HPO4)4] (I) and [C8N4H26][Zn6(PO4)4(HPO4)2] (II) have been synthesized employing solvothermal reactions in the presence of N,N(3-bisaminopropyl)-1,2-ethylenediamine. The structure of I consists of ZnO4 and HPO4 tetrahedra, forming four-membered rings, which are connected edgewise giving rise to the one-dimensional ladders. HPO4 units also hang from the Zn center into the inter-ladder spaces along with the organic amine molecules. In II, the connectivity between ZnO4, PO4 and HPO4 tetrahedral units gives rise to a two-dimensional layered structure with eight-membered apertures. The amine molecule occupies the center of these apertures and interacts with the layer through hydrogen bonds. The formation of one-dimensional tube-like structure in II is noteworthy.  相似文献   

11.
In addition to reversible O2 binding, respiratory proteins of the globin family, hemoglobin (Hb) and myoglobin (Mb), participate in redox reactions with various metal complexes, including biologically significant ones, such as those of copper and iron. HbO2 and MbO2 are present in cells in large amounts and, as redox agents, can contribute to maintaining cell redox state and resisting oxidative stress. Divalent copper complexes with high redox potentials (E 0, 200-600 mV) and high stability constants, such as [Cu(phen)2]2+, [Cu(dmphen)2]2+, and CuDTA oxidize ferrous heme proteins by the simple outer-sphere electron transfer mechanism through overlapping π-orbitals of the heme and the copper complex. Weaker oxidants, such as Cu2+, CuEDTA, CuNTA, CuCit, CuATP, and CuHis (E 0≤ 100-150 mV) react with HbO2 and MbO2 through preliminary binding to the protein with substitution of the metal ligands with protein groups and subsequent intramolecular electron transfer in the complex (the site-specific outer-sphere electron transfer mechanism). Oxidation of HbO2 and MbO2 by potassium ferricyanide and Fe(3) complexes with NTA, EDTA, CDTA, ATP, 2,3-DPG, citrate, and pyrophosphate PPi proceeds mainly through the simple outer-sphere electron transfer mechanism via the exposed heme edge. According to Marcus theory, the rate of this reaction correlates with the difference in redox potentials of the reagents and their self-exchange rates. For charged reagents, the reaction may be preceded by their nonspecific binding to the protein due to electrostatic interactions. The reactions of LbO2 with carboxylate Fe complexes, unlike its reactions with ferricyanide, occur via the site-specific outer-sphere electron transfer mechanism, even though the same reagents oxidize structurally similar MbO2 and cytochrome b 5 via the simple outer-sphere electron transfer mechanism. Of particular biological interest is HbO2 and MbO2 transformation into met-forms in the presence of small amounts of metal ions or complexes (catalysis), which, until recently, had been demonstrated only for copper compounds with intermediate redox potentials. The main contribution to the reaction rate comes from copper binding to the “inner” histidines, His97 (0.66 nm from the heme) that forms a hydrogen bond with the heme propionate COO group, and the distal His64. The affinity of both histidines for copper is much lower than that of the surface histidines residues, and they are inaccessible for modification with chemical reagents. However, it was found recently that the high-potential Fe(3) complex, potassium ferricyanide (400 mV), at a 5 to 20% of molar protein concentration can be an efficient catalyst of MbO2 oxidation into metMb. The catalytic process includes binding of ferrocyanide anion in the region of the His119 residue due to the presence there of a large positive local electrostatic potential and existence of a “pocket” formed by Lys16, Ala19, Asp20, and Arg118 that is sufficient to accommodate [Fe(CN)6]4–. Fast, proton-assisted reoxidation of the bound ferrocyanide by oxygen (which is required for completion of the catalytic cycle), unlike slow [Fe(CN)6]4– oxidation in solution, is provided by the optimal location of neighboring protonated His113 and His116, as it occurs in the enzyme active site.  相似文献   

12.
Since there is evidence that oxalyl thiolesters (RSCOCOO) are present in animal cells, and possibly may participate in the control of metabolism, the present study was undertaken to characterize their reactivity with nucleophiles so that one could gain a better understanding of how they might be affecting the activities of enzymes. At 25°C and neutral pH, N-acetyl-S-oxalyl-2-aminoethanethiol (NAC-S-Ox) reacts rapidly with cysteamine (2-aminoethanethiol) to give N-acetylcysteamine and N-oxalylcysteamine. Under similar conditions, other aminothiols, such as cysteine, homocysteine, penicillamine, and cysteine ethyl ester, also react rapidly with NAC-S-Ox, but non-thiol-containing amines, such as alanine, alanine ethyl ester, glycine, and S-methylcysteine, react more than four orders of magnitude less rapidly. The aminothiol reactions apparently proceed by rate-determining oxalyl transfer to the thiol followed by a rapid intramolecular S- to N-oxalyl migration. The reactions follow second-order kinetics with the thiolate anion being the reactive nucleophile. At 25°C and ionic strength 1.0 , kN, defined in the equation, rate = kN[RS][NAC-S-Ox], has the following values ( −1 s−1) for the anion of the reacting thiol: cysteamine, 170; cysteine, 260; cysteine ethyl ester, 76; homocysteine, 380. Rate data for the reaction of NAC-S-Ox with hydroxylamine, imidazole, hydroperoxide, and hydroxide were also obtained. The reaction of S-oxalyl-p-thiocresol with thiol anions under the same conditions gives the following values for kN ( −1 s−1 × 10−3): glutathione, 5.6; N-acetylcysteamine, 3.7; pantetheine, 4.8; 8-mercaptooctanoic acid, 4.5; 6-mercaptooctanoic acid, 1.0; dihydrolipoic acid, 8.2. These results indicate that oxalyl transfers from oxalyl thiolesters to thiol anions occur more than two orders of magnitude more rapidly than corresponding acetyl transfers, and that under physiological conditions any in vivo oxalyl thiolester would equilibrate within minutes with virtually every thiol in the cell, including those attached to enzymes. Consequently, it is proposed that one mechanism by which oxalyl thiolesters may function in vivo to alter the catalytic activities of enzymes is to covalently modify enzymic thiols by acylation with an oxalyl group.  相似文献   

13.
In the coupling of ATP pyrophosphorolysis to Ca2+ transport in beef heart mitochondria, for each molecule of ATP cleaved, one proton is released and one Ca2+ is transported into the interior space. With the use of tritium labelled ATP, it could be shown that ATP is pyrophosphorylyzed into a species equivalent to Pi that moves inward, and a species equivalent to ADP that is extruded into the aqueous space on the exterior of the mitochondrion. The species equivalent to Pi has been identified as a negatively charged form of Pi (PO?) and the species equivalent to ADP as a positively charged form (ADP+). The inward flow of PO? is coupled to the inward flow of Ca2+ in 1:1 stoichiometry—a token that Ca2+ must enter in the form of Ca2+A?, where A? is a monovalent anion. During ATP pyrophosphorolysis protons are released on the I side and taken up on the M side—one proton for each molecule of ATP cleaved. The alkalinization of the matrix space leads to the deposition of Ca3(PO4)2 and to the extrusion of the two species released by this deposition (Pi, K+). Two thirds of the PO? is trapped as Ca3(PO4)2 in the matrix space and one third is extruded into the external space. The extrusion of K+ provides a mechanism by which protons can be continuously brought into the matrix space to sustain a high rate of coupled pyrophosphorolysis of ATP. The coupling pattern for Ca2+ transport driven by ATP pyrophosphorolysis is identical with the corresponding pattern for Ca2+ transport driven by electron transfer. This identity is suggestive that coupling mediated by the Fo-F1 system and coupling mediated by electron transfer complexes are mechanistically indistinguishable.  相似文献   

14.
C.A. Wraight 《BBA》1979,548(2):309-327
The photoreduction of ubiquinone in the electron acceptor complex (Q1Q11) of photosynthetic reaction centers from Rhodopseudomonas sphaeroides, R26, was studied in a series of short, saturating flashes. The specific involvement of H+ in the reduction was revealed by the pH dependence of the electron transfer events and by net H+ binding during the formation of ubiquinol, which requires two turnovers of the photochemical act. On the first flash Q11 receives an electron via Q1 to form a stable ubisemiquinone anion (Q??11); the second flash generates Q??1. At low pH the two semiquinones rapidly disproportionate with the uptake of 2 H+, to produce Q11H2. This yields out-of-phase binary oscillations for the formation of anionic semiquinone and for H+ uptake. Above pH 6 there is a progressive increase in H+ binding on the first flash and an equivalent decrease in binding on the second flash until, at about pH 9.5, the extent of H+ binding is the same on all flashes. The semiquinone oscillations, however, are undiminished up to pH 9. It is suggested that a non-chromophoric, acid-base group undergoes a pK shift in response to the appearance of the anionic semiquinone and that this group is the site of protonation on the first flash. The acid-base group, which may be in the reaction center protein, appears to be subsequently involved in the protonation events leading to fully reduced ubiquinol. The other proton in the two electron reduction of ubiquinone is always taken up on the second flash and is bound directly to Q??11. At pH values above 8.0, it is rate limiting for the disproportionation and the kinetics, which are diffusion controlled, are properly responsive to the prevailing pH. Below pH 8, however, a further step in the reaction mechanism was shown to be rate limiting for both H+ binding electron transfer following the second flash.  相似文献   

15.
(NIn)‐Formyl protective group of tryptophan has been introduced as a base/nucleophile‐labile protective group. It has long been known that a free ‐amino group of the peptide can serve as a nucleophile: an irreversible formyl NIn → NH2 transfer is consistently observed when deformylation is performed last on an otherwise deprotected peptide that possesses free ‐amino group. Obviously, this particular side reaction should be expected any time free amino group is exposed to Trp(For), but, at the best of our knowledge, has never been reported in the course of Boc‐SPPS. In the present communication, we describe a set of appropriately designed model experiments that permitted to detect the title side reaction both in solution and in solid‐phase reactions. We observed intermolecular formyl group transfer with a model compound, Trp(For)‐NH2. Importantly, we also observed this migration on solid support with the rate roughly estimated to be up to 1% of residues per minute. We also observed that the formyl‐group transfer reaction occurred in a sequence‐dependent manner and was suppressed to a non‐detectable level using ‘in situ neutralization’ technique. Because this side reaction is sequence dependent, there might be situations when the rate of the formation of Nα‐formyl termination by‐products is significant. In other cases, the Nα‐For truncated by‐products would not contaminate the final peptide significantly but still could be a source of microheterogeneity. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
Oxidation of the ferrous derivatives of earthworm erythrocruorin with potassium ferricyanide at pH 7 gives rise to met-erythrocruorin with the characteristic spectral properties of aquo-met-haem proteins. Met-erythrocruorin maintains the same overall conformation (s20,w, α-helicalcontent) as the native protein. However, its stability is limited to a very narrow pH range around neutrality; outside this range it is converted rapidly and irreversibly to another spectral species (hemichrome). The erythrocruorin-hemichrome undergoes a reversible pH-dependent transition (pK = 9.2) which is accompanied by a decrease in α-helical structure.On reduction the met-form and the hemichrome yield the deoxy and oxy derivatives. Hemichrome formation is accompanied by a drastic change of the quaternary structure; thus the sedimentation coefficient drops from 60 S to ~4 S and the α-helical content decreases. By addition of ligands (CN?, N3?) or by reduction to the ferrous form, the hemichrome reassociates into 10 S subunits.  相似文献   

17.
The solubilities of H2S in ionic liquids, 1-butyl-3-methylimidazolium tetrafluoroborate ([C4mim+][BF4?]), 1-butyl-3-methylimidazolium hexaflorophosphate ([C4mim+][PF6?]) and -butyl-3-methylimidazolium tetrafluoroborate bis(trifluoromethanesulphonyl)amide ([C4mim+][Tf2N?]) are predicted using isothermal–isobaric Gibbs ensemble Monte Carlo method (NPT-GEMC) at temperatures ranging from 333 to 453 K and pressure up to 20 bar. The low pressure points (up to 3 bar) of the absorption isotherms are fitted to a straight line to get a rough estimation of the Henry’s law constants. The van’t Hoff relationship is used to estimate the partial molar enthalpy of the absorption. The obtained results are in a good agreement with Jou and Mather [18], and Jalili et al. [13]. When comparing the solubility in ILs, it is found that H2S solubility is highest for [C4mim+][Tf2N?], followed by [C4mim+][PF6?]. The lowest solubility is observed in [C4mim+][BF4?]. The highest solubility in [C4mim+][Tf2N?] is consistent with Jalili et al. [13]. However, their results indicate slightly higher solubility in [C4mim+][BF4?] than in [C4mim+][PF6?], and do not agree with the predictions. Upon absorption, the molar volumes of the mixtures decrease linearly, showing only small changes in volume. The effect of H2S absorption on ILs is further studied by calculating the radial distribution functions between the ions. The results indicate that the solute molecules accommodate themselves in the cavities without significantly disturbing the ionic arrangement of the ions, similar to CO2 absorption in ILs. The spatial distribution functions show similar spatial distribution for H2S around cation in all of the studied ILs, whereas the distribution around anion depends on the shape and flexibility of the anion. The mechanism of H2S absorption is studied by computing the van der Waals (VDW) and electrostatic (ELEC) energies. It is observed that the solubility of H2S in the studied ILs is primarily controlled by VDW interaction. When comparing the interaction of H2S with the ions, it is found that solute molecules interact with cations mainly due to VDW interaction. Both VDW and ELEC energies contribute in the interaction between H2S and anions.  相似文献   

18.
Magnetic circular dichroism (MCD) spectroscopy has been used to explore the connection between optical spectra and the high spin population of several hemoglobins under various conditions. It is found that the effectiveness of IHP in inducing spectral changes can be markedly affected by solvent. For example, the IHP-induced spectral changes in the visible region for nitritomethemoglobin-A in mixed buffer solvent systems (glycerol or polyethylene glycol (PEG), mw 190–210) are more than double those observed in aqueous buffers. We estimate that IHP induces a mix of R/T forms in bis-tris phosphate buffers, for NO2?metHb that is only about 50% T form. While PEG and glycerol both lead to enhanced IHP-induced spectral differences, they behave differently in two aspects. PEG shifts the visible MCD and absorption spectra of F?metHb-A. supposedly already biased towards the T form by ligand, in the same direction that IHP does. PEG also maximizes the spin state changes with IHP for three R form hemoglobins and N3?metHb-A, and so appears to stabilize the T form in all cases. Glycerol does not. In addition, the apparent binding constant for NO2? to H2OmetHb-A differs between these two solvents. Comparison of the data from several hemoglobins leads to the conclusion that the changes in spin state distributions induced by IHP correlate well with quarternary structure for a given hemoglobin. An analogous correlation amongst various proteins between initial spin state distribution (IHP) absent) and quarternary structure is not found.  相似文献   

19.
This study describes the relationships between dinitrogen (N2) fixation, dihydrogen (H2) production, and electron transport associated with photosynthesis and respiration in the marine cyanobacterium Trichodesmium erythraeum Ehrenb. strain IMS101. The ratio of H2 produced:N2 fixed (H2:N2) was controlled by the light intensity and by the light spectral composition and was affected by the growth irradiance level. For Trichodesmium cells grown at 50 μmol photons · m?2 · s?1, the rate of N2 fixation, as measured by acetylene reduction, saturated at light intensities of 200 μmol photons · m?2 · s?1. In contrast, net H2 production continued to increase with light levels up to 1,000 μmol photons · m?2 · s?1. The H2:N2 ratios increased monotonically with irradiance, and the variable fluorescence measured using a fast repetition rate fluorometer (FRRF) revealed that this increase was accompanied by a progressive reduction of the plastoquinone (PQ) pool. Additions of 2,5‐dibromo‐3‐methyl‐6‐isopropyl‐p‐benzoquinone (DBMIB), an inhibitor of electron transport from PQ pool to PSI, diminished both N2 fixation and net H2 production, while the H2:N2 ratio increased with increasing level of PQ pool reduction. In the presence of 3‐(3,4‐dichlorophenyl)‐1,1‐dimethylurea (DCMU), nitrogenase activity declined but could be prolonged by increasing the light intensity and by removing the oxygen supply. These results on the coupling of N2 fixation and H2 cycling in Trichodesmium indicate how light intensity and light spectral quality of the open ocean can influence the H2:N2 ratio and modulate net H2 production.  相似文献   

20.
New efficient catalysts based on electrophilic N‐fluoro quaternary ammonium salts are reported for catalytic allylation of (E)‐N,1‐diphenylmethanimine. The chiral version of the catalyst based on cinchonidine (F‐CD‐BF4) shows high catalytic activity with approximately 94% ee and TOF (>800 h?1). The F‐CD‐BF4 is prepared from cinchonidine and Selectfluor by one‐step transfer fluorination.  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