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.
The stability constants of the 1:1 complexes formed between Cu(Arm)2+, where Arm = 2,2′-bipyridyl or 1,10-phenanthroline, and methyl phosphate, CH3OPO32−, or hydrogen phosphate, HOPO32−, were determined by potentiometric pH titration in aqueous solution (25°C; l = 0.1 M, NaNO3). On the basis of previously established log K versus pKa straight-line plots (D. Chen et al., J. Chem. Soc., Dalton Trans. (1993) 1537–1546) for the complexes of simple phosphate monoesters and phosphonate derivatives, R-PO32−, where R is a non-coordinating residue, it is shown that the stabilities of the Cu(Arm) (CH3OPO3) complexes are solely determined by the basicity of the -PO32− residue. In contrast, the Cu(Arm) (HOPO3) complexes are slightly more stable (on average by 0.15 log unit) than expected on the basicity of HPO42−; this is possibly due to a more effective solvation including hydrogen bonding, an interaction not possible with coordinated CH3OPO32− species. Regarding biological systems the observation that HOPO32− is somewhat favored over R-PO32− species in metal ion interactions is meaningful.  相似文献   

2.
Cuaq+ forms stable complexes with carbon monoxide in aqueous solutions. Furthermore it reacts very fast with aliphatic radicals. The reaction of Cu(CO)maq+ with methyl radicals, CH3 was studied using the pulse-radiolysis technique. The results point out that methyl radicals react with Cu(CO)aq+ to form an unstable intermediate with a CuII-C σ bond identified as (CO)CuII-CH3+, k = (1.1±0.2) × 109 M−1 s−1. This intermediate has a strong LMCT charge transfer band (λmax = 385 nm, max = 2500 M−1 cm−1) which is similar to the absorption bands of other transient complexes with CuII-alkyl σ bonds. The coordinated carbon monoxide in (CO)CuII-CH3+ inserts into the copper—carbon bond (or rather the coordinated methyl migrates to the coordinated carbon monoxide ligand) at a rate of (3.0±0.8) × 102 s−1 to form the copperacetyl complex (CO)mCuII-C(CH3)=O+max = 480 nm, max = 2100 M−1 cm−1). The rate of formation of (CO)CuII-CH3+ and of the insertion reaction are pH independent. The complex (CO)mCuII-C(CH3)=O+ is also unstable and decomposes heterolytically to yield acetaldehyde and Cuaq2+ as the final stable products. This reaction is slightly pH dependent. The same reactivity pattern has been observed for the Cu(COnaq+ complexes (n = 2 or 3). The results clearly point out that CO remains coordinated to transient complexes of the type CuII-alkyl.  相似文献   

3.
The preparation and reaction chemistry of 1,3- and 1,2-diene and related complexes derived from metal carbonyl containing anions and allenic electrophiles are addressed. The preparation of some CpFe(CO)2 η1-diene complexes and their conversion into CpFe(CO) η3-diene complexes is presented followed by reactions of CpMo(CO)3, CpW(CO)3 and CpMo(CO)2PR3 anions with allenic electrophiles which produce metal complexed cyclobutenones (via CO and alkene insertions from the initially formed product) and 1,2-diene complexes, respectively. Lastly, the reactions of PPh3(CO)3Co anions with allenic electrophiles are outlined which result in several different coordination geometries depending on the reaction conditions used.  相似文献   

4.
Due to contradictions in the literature we have redetermined the acid-base properties of riboflavin (=RiFl; vitamin B2), i.e. 7,8-dimethyl-10-ribityl-isoalloxazine, and of flavin mononucleotide (FMN2−), also known as riboflavin 5′-phosphate, via potentiometric pH titrations (I = 0.1 M, NaNO3; 25 °C). In contrast to various claims, the isoalloxazine ring cannot be protonated at pH > 1, a result in agreement with an early study (pKa = −0.2; L. Michaelis, M.P. Schubert and C.V. Smythe, J. Biol. Chem., 116 (1936) 587–607); deprotonation of the ring system occurs in both compounds with pKa 10. The pKa value of 0.7 determined for the deprotonation of H2(FMN) must be attributed to the release of the first proton from the fully protonated phosphate group; its second proton is released with pKa = 6.18 in agreement with the acidity constants of various other monoprotonated monophosphate esters. The stability constants of the 1:1 complexes formed between Mg2+, Ca2+, Sr2+, Ba2+, Mn2+, Co2+, Ni2+, Cu2+, Zn2+ or Cd2+ (---M2+) and FMN2− were determined by potentiometric pH titrations in aqueous solution (I = 0.1 M, NaNO3; 25 °C). The log stability constants of all these M(FMN) complexes are about 0.2 log units higher than expected from the basicity of the phosphate group. This slight stability increase cannot be attributed to the formation of a seven-membered chelate involving the ribit-hydroxy group at C-4′ as the stability constants for the M2+ 1:1 complexes of glycerol 1-phosphate (G1P2−) demonstrate: G1P2− contains the same structural unit which would also allow in this case the formation of the mentioned seven-membered chelate; however, the stability of the M(G1P) complexes is solely determined by the basicity of the phosphate group. Hence, in agreement with earlier conclusions (J. Bidwell, J. Thomas and J. Stuehr, J. Am. Chem. Soc., 108 (1986) 820–825) regarding Ni(FMN) one must conclude that the slight stability increase of the M(FMN) complexes has to be attributed to the isoalloxazine ring. The equality of the stability increase of the complexes for all the mentioned ten metal ions precludes its attribution to an interaction with an N site and makes a specific interaction with an O site also somewhat unlikely. In addition, carbonyl oxygens appear as not very favorable for the formation of macrochelates by a further interaction with already phosphate-coordinated metal ions. Therefore, we propose that the slight but significant stability increase originates from M(FMN) species (with a formation degree of about 30%) in which the hydrophobic flavin residue is close to the metal ion, thereby lowering the ‘effective’ dielectric constant in the microenvironment of the metal ion and thus indirectly promoting the −PO32−/M2+ interaction.  相似文献   

5.
The inhibitions by Ni2+ and F ions and by acetohydroxamic acid of jack bean urease covalently immobilized on chitosan membrane was studied (pH 7.0, 25°C) and compared with those of the native enzyme. The reaction progress curves of the immobilized urease-catalyzed hydrolysis of urea were recorded in the absence and presence of the inhibitors. They revealed that the inhibitions are of the competitive slow-binding type similar to those of native urease. The immobilization weakened the inhibitory effect of the inhibitors on urease as measured by the inhibition constants Ki*. The increase in their values: 17.9-fold for Ni2+, 26.5-fold for F and 1.7-fold for acetohydroxamic acid, was accounted for by environmental effects generated by heterogeneity of the urease–chitosan system: (1) mass transfer limitations imposed on substrate and reaction product in the external solution, and (2) the increase in local pH on the membrane produced by both the enzymatic reaction and the electric charge of the support. By relating the KM/Ki* ratio to the electrostatic potential of chitosan it was found that while the reduced Ni2+ inhibition is mainly brought about by the potential, inhibition by acetohydroxamic acid is independent of the potential, and the acid inhibits urease in its non-ionic form. The reduction in F inhibition was ascribed to the increased pH in the local environment of the immobilized enzyme.  相似文献   

6.
J. Butler  G.G. Jayson  A.J. Swallow 《BBA》1975,408(3):215-222

1. 1. The superoxide anion radical (O2) reacts with ferricytochrome c to form ferrocytochrome c. No intermediate complexes are observable. No reaction could be detected between O2 and ferrocytochrome c.

2. 2. At 20 °C the rate constant for the reaction at pH 4.7 to 6.7 is 1.4 · 106 M−1 · s−1 and as the pH increases above 6.7 the rate constant steadily decreases. The dependence on pH is the same for tuna heart and horse heart cytochrome c. No reaction could be demonstrated between O2 and the form of cytochrome c which exists above pH ≈ 9.2. The dependence of the rate constant on pH can be explained if cytochrome c has pKs of 7.45 and 9.2, and O2 reacts with the form present below pH 7.45 with k = 1.4 · 106 M−1 · s−1, the form above pH 7.45 with k = 3.0 · 105 M−1 · s−1, and the form present above pH 9.2 with k = 0.

3. 3. The reaction has an activation energy of 20 kJ mol−1 and an enthalpy of activation at 25 °C of 18 kJ mol−1 both above and below pH 7.45. It is suggested that O2 may reduce cytochrome c through a track composed of aromatic amino acids, and that little protein rearrangement is required for the formation of the activated complex.

4. 4. No reduction of ferricytochrome c by HO2 radicals could be demonstrated at pH 1.2–6.2 but at pH 5.3, HO2 radicals oxidize ferrocytochrome c with a rate constant of about 5 · 105–5 · 106 M−1 · s−1

.  相似文献   


7.
Reactions of cis-diamminedichloroplatinum(II) with phosphonoformic acid (PFA), phosphonoacetic acid (PAA), and methylenediphosphonic acid (MDP) yield various phosphonatoplatinum(II) chelates which were characterized by phosphorus-31 NMR spectroscopy. The P-31 resonances for the chelates appear at 6–12 ppm downfield as compared to the uncomplexed ligands. All complexes exhibit monoprotic acidic behavior in the pH range 2–10. The chemical shift-pH profiles yielded acidity constants, 1.0 × 10−4, 1.5 × 10−4, and 1.3 × 10−6 M−1, for the PFA, PAA, and MDP chelates. In addition to the monomeric chelate, MDP formed a bridged diplatinum(II,II) complex when it reacted with cis-Pt (NH3)2(H2O)22+. The P-31 resonance for this binuclear complex appears at 22 ppm downfield from the unreacted ligand.

Rate data for the complexation reactions of the phosphonate ligands with the dichloroplatinum complex are consistent with a mechanism in which a monodentate complex is formed initially through rate-limiting aquation process of the platinum complex, followed by a rapid chelation. For the PFA and PAA complexes, initial binding sites are the carboxylato oxygens. Implications of the various binding modes of the phosphonates in relationship to their antiviral activities are discussed.  相似文献   


8.
Crab shell particles were used as a biosorbent to remove lead from aqueous solutions. The equilibrium isotherm showed that crab shell particles took up lead to the extent of 1300 mg Pb g−1 crab shell. The optimum pH range for maximum lead removal was increased to 5·5–11·0 compared to the shell-free control pH of 8·5–11·0. pH values of solutions with crab shell material added were increased spontaneously to about 10 as a result of the CaCO3 present, which formed complexes with lead according to pH. Electron spectroscopy, Fourier transform infrared spectrometry, scanning electron microscopy and X-ray diffraction results confirmed that -NHCOCH3 and CO32 were involved in binding of lead. In addition, the removal of lead occurred mainly through dissolution of CaCO3 followed by precipitation of Pb3(CO3)2(OH)2 and PbCO3 near the surface of crab shell. Micro precipitates formed were then adsorbed to the chitin on the surface of the crab shell particles.  相似文献   

9.
The perchlorate (ClO4)-respiring organism, strain perc1ace, can grow using nitrate (NO3) as a terminal electron acceptor. In resting cell suspensions, NO3 grown cells reduced ClO4, and ClO4 grown cells reduced NO3. Activity assays showed that nitrate reductase (NR) activity was 1.31 μmol min−1 (mg protein)−1 in ClO4 grown cells, and perchlorate reductase (PR) activity was 4.24 μmol min−1 (mg protein)−1 in NO3 grown cells. PR activity was detected within the periplasmic space, with activities as high as 14 μmol min−1 (mg protein)−1. The NR had a pH optimum of 9.0 while the PR had an optimum of 8.0. This study suggests that separate terminal reductases are present in strain perclace to reduce NO3 and ClO4.  相似文献   

10.
The effect of pH of electrolyte solution on the interfacial tension of lipid membrane formed of phosphatidylcholine (PC, lecithin)–phosphatidylserine (PS) system was studied. In this article, three models describing the H+ and OH ions adsorption in the bilayer lipid surface are presented. In Model I and Model II, the surface is continuous with uniformly distributed functional groups constituting the centres of H+ and OH ions adsorption while in the other the surface is built of lipid molecules, free or with attached H+ and OH ions. In these models contribution of the individual lipid molecule forms to interfacial tension of the bilayer were assumed to be additive. In Model III the adsorption of the H+ and OH ions at the PC–PS bilayer surface was described in terms of the Gibbs isotherm. Theoretical equations are derived to describe this dependence in the whole pH range.  相似文献   

11.
Electron self-exchange in solutions of the ‘blue’ copper protein plastocyanin is catalysed by the redox-inert multivalent cations Mg2+ or Co(NH3)3+6. Measurements of specific 1H-NMR line broadening with 50% reduced solutions in the presence of these cations show that electron exchange proceeds through encounters of cation-protein complexes which dissociate at high ionic strength. In the presence of 8mM (5 equivalents/total protein) Co(NH3)3+6, with 10 mM cacodylate (pH*6.0) as background electrolyte, the bimolecular rate constant at 25°C is 7 × 104 M−1·s−1. For comparison, the ‘electrostatically screened’ rate constant measured in 0.1 M KCl in the absence of added multivalent cations is ˜ 4 × 103 M1·s−1.

Plastocyanin Electron self-exchange NMR Protein-protein interaction Multivalent cation Blue copper protein  相似文献   


12.
Rate constants determined by the stopped-flow method for four protein-protein reactions at 25°C, pH's in the range 5.8–7.5. I = 0.10 M (NaCI), are as follows: cytochrome c(II) with plastocyanin, PCu(II). 1.5 × 106 M−1 sec−1, pH 7.6; high-potential iron-sulfur protein (Hipip) with PCu(II), 3.7 × 105 M−1sec−1. pH 5.8; cytochrome c(II) with azurin, ACu(ll). 6.4 × 103 M−1sec−1, pH 6.1; Hipip with ACu(II), 2.2 × 105 M−1sec−1, pH 5.8. Activation parameters have been determined for all four reactions; they indicate higher enthalpy requirements and less negative entropy requirements for the PCu(II) as opposed to ACu(II) reactions. Equilibrium constants K for association prior to electron transfer are < 150 M−1 for the cytochrome c(II) reduction of PCu(II) (estimated charges 8 + and 9-,respectively), and < 300 M−1 for the other reactions, indicating no favorable interactions. Rate constants have been analyzed in terms of the simple Marcus theory, which has previously given an excellent fit to thirteen protein-protein reactions considered by Wherland and Pecht. No similar correlation exists in the present studies, and calculated rate constants differ by orders of magnitude from experimentally determined values.  相似文献   

13.
The two uncharged compounds 25,26,27,28-(2-N,N-di methyldithiocarbamoylethoxy)calix[4]arene (1) and 25,26,27,28- (2-methylthioethoxy)calix[4]arene (2) are effective extractants for transferring Hg(II), Ag(I), Pd(II) and Au(III) from aqueous solution into chloroform. The electronic absorption spectra of 1 and 2 show additional bands at long wavelength upon complexation with AuCl4, PdCl42− and PdBr42−, and analogous bands for Hg2+ and Ag+ with 1. For 1 these new bands are considered to be either of the charge transfer type or transitions within the C=S moiety. These new bands for the complexes with 2 are assigned to LMCT transitions of the S → M type. These spectral features are used to obtain information about the solution structures of the complexes that are formed between these metal ions and both 1 and 2.  相似文献   

14.
Resonance Raman measurements have been performed with solutions of iodine-complexed synthetic amyloses (DP 25–200), malto-oligomers (DP 3–18, and -cylodextrin. Interest was focused on the minimum chain length for helical complex formation and a possible preferred length for the polyiodine chain. Four fundamental vibrations are observed at 164, 112, 52 and 24 cm−1. The 112 cm−1 Raman line was shown to arise both from free I3 (enhanced at 363.8 nm excitation) and from bound iodine (relatively most intense at 457.9 nm excitation). The main signal of complexed iodine at 164 cm−1is enhanced at an excitation wavelength close to the long wavelength absorption maximum. This signal is observed firt with malto-octaose and -cyclodextrin. The less intense signals at 52 and 24−1 are only detected at DP 15 and higher. Raman spectra give no evidence for a preferred length of the polyiodine chain. Significantly identical Raman spectra are obtained when using different molar ratios of I2/KI solution or I2 solution initially free of I ions. The results are discussed in view of previous assignments of the Raman lines to I2, I3/I2, and I5 subunits. Our findings are incompatible with I3 units as the only bound species. They are compatible with both I3/I2 and I3 subunits under certain conditions. In the case of I2 solution used for complexation we favour the polyiodine chain model proposed previously by Cramer35,36. The I3 ions formed could function mainly as chain initiators, as has been suggested by Cesàro and Brant30.  相似文献   

15.
1. The alteration of the Ca2+ requirements of the ATPase activity of fibrils from rabbits and crabs at varying ionic strength, pH and concentration of MgATP (i.e. MgATP2− + MgHATP) was investigated.

2. Under physiological conditions, it was found that the ATPase activity of rabbit and crab fibrils after an initial increase decreased steeply when the Ca2+ concentration is raised above 1×10−4 M. This is a primary effect of the over-optimal Ca2+ concentration and not a secondary one caused by the influence of accompanying ions.

3. The Ca2+ requirements for ATP splitting by rabbit fibrils remain constant at an ionic strength from 0.1 to 0.2 and for a MgATP concentration in the range from 0.5 to 10 mM. At I = 0.05 it is about 5 times smaller than at 0.1. When the pH is decreased from 8 to 7, the Ca2+ requirements are increased some 10 times but only 3 times when the pH is varied between 7 and 6.

4. In crab fibrils, there is no alteration of the Ca2+ requirements when the ionic strength is varied between 0.05 and 0.2, but a reduction of the pH from 8.0 to 6.0 raises the Ca2+ requirements for half activation and for threshold by a factor of 10. Changing the MgATP concentration increases the Ca2+ requirements only in the range from 1 to 5 mM, while the concentration required in 0.5 mM is identical with that at 1 mM, and 10 mM corresponds to 5 mM.

5. It can be deduced from the experimental results that at a pH above 6.0 maximal activation is always obtained if the Ca2+ concentration is 5×10−5 M. By contrast, relaxation is only achieved when the Ca2+ concentration is below 1×10−7 M for pH 7.0 and I > 0.1 or below 1×10−8 for pH > 7.0 or I < 0.1.

6. To achieve complete relaxation, an ethyleneglycoldiaminotetraacetate (EGTA) concentration of 1 mM is sufficient, even when there is a large degree of contamination by Ca2+ as long as the pH stays above 6.5.  相似文献   


16.
The kinetics of the anation reactions of [M(RNH2)5H2O]3+ (M = Rh, R = H, Me, Et, Pr; M = Cr, R = H, Me, Pr) with several ligands (H3PO4/H2PO4, H3PO3/H2PO3, CF3COO, Br, Cl, SCN) have been studied at different temperatures and acidities at I = 1.0 M (LiClO4. Results obtained for the anation rate constants and thermal activation parameters are compared with the previously published data for R = H, in order to establish the effects of the amine substituents in the reaction mechanism proposed for the substitution reactions of these complexes. The results obtained are interpreted on the basis of a mechanism where the bond formation process is more important in the substitution on M = Cr complexes than in that of the M = Rh complexes, as already pointed out for the published ΔΛ values for the water exchange on these systems. A simple Langford-Gray classification becomes inadequate to describe these situations where the increase of the steric demand of the amine substituents shifta the Ia-Id classification to the Id side, although no dramatic changes in the reaction mechanism are found. It is concluded that a More O'Ferall ‘continuous’ type of approach to the mechanism classification of the substitution reactions is much more useful in this case.  相似文献   

17.
A new functional macrocyclic ligand, 2,4-dinitrophenylcyclen (= 1-(2,4-dinitrophenyl)-1,4,7,10-tetraazacyclododecane), has been synthesized and isolated as its trihydrochloric acid salt (L·3HCl). The protonation constants (log Kn) for three secondary nitrogens of L were determined by potentiometric pH titration to be 10.10, 7.33 and <2 with I = 0.10 (NaNO3) at 25°C. The 2,4-dinitrophenylaniline chromophore was proven to be a good reporter signaling proton- and metal-binding events in the macrocyclic cavity. The UV absorption band (λmax 370 nm, 8200) of the 2,4-dinitrophenylaniline moiety at pH ≥ 9 becomes quenched as pH is lowered (to pH 3.1, where the major species is L·2H+), due to the strong protonation effect extended to the aniline moiety within the macrocyclic cavity. This is in sharp contrast to the pH-independent UV absorption (λmax 390 nm, 14 000) of a reference compound, N,N-diethyl-2,4-dinitroaniline. The UV absorption band of L is shifted to lower wavelengths with Zn2+max 320 nm), Cd2+max 316 nm) and Pb2+max 317 nm), while it almost disappears with Cu2+ and Ni2+. The 1:1 Zn2+ and Cu2+ complexes with L were isolated and characterized. The Zn2+ complex recognizes 1-methylthymine anion (MT) in aqueous solution at physiological pH to yield a stable ternary complex ZnL-MT. The X-ray crystal structure of ZnL-MT showed that Zn2+ is four-coordinate with three secondary nitrogens of L and the deprotonated imide anion that is cofacial to the 2,4-dinitrophenyl ring.  相似文献   

18.
[RuII(Me2edda)(H2O)2] (1), Me2edda2− = N,N′-dimethylethylenediaminediacetate, exhibits a sterically-controlled molecular recognition in forming η2 and η4 olefin complexes. 1 exists with an N2O2 in-plane set of chelate donors and axial H2O ligands. The two CH3 functionalities of Me2edda2− are poised above and below the N2O2 plane of the glycinato rings. Studies herein of the 2,2′-bipyridine complex, [RuII(Me2edda)(bpy)], with bidentate bpy chelation as established via 1H NMR and electrochemical methods show 1 to be ligated in the S,S configuration with the glycinato rings in-plane as a cis-O form. 1 is sterically discriminating in forming η2 complexes with smaller olefins (ethylene, 2-propene, cis-2-butene, methyl vinyl ketone and 3-cyclohexene-1-methanol), but rejects larger decorated ring structures and branched olefins (1,2-dimethyluracil, cyclohexene-1-one 2-methyl-2-propene). η2 complexes of 1 have characteristic RuII/III DPP waves near 0.55 V which vary slightly with olefin structure. Potentially bidendate dienes (1,3-butadiene, 1,3-cyclohexadiene and 2,5-norbornadiene (nbd) form η4 complexes as shown by RuII/III waves between 0.94 and 1.30 V, indicate of a highly stabilized RuII center by π-backboning. An η2η4 ‘equilibrium’ with apparent K = 22 at 25 °C is observed for nbd coordinated to 1. (The η2 and η4 distribution may be a kinetic one and not a thermodynamic one). To allow formation of the cis η4 complexes, 1 must undergo a shift of one or both glycinato donors from the N2O2 plane into the axial site away from the dimethyl functionalities. η4 chelation by 1,3-butadiene has been confirmed by 1H NMR spectral assignments of two [RuII(Me2edda)] isomers, one in the axial rans-O glycinato configuration, e.g. 1,3-butadiene is bidentate in the original N2O2 plane and a second unsymmetrical glycinato arrangement with in-plane and axial glycinato as well as in-plane and axial η4-1,3-butadiene coordination. [RuII(hedta)(H2O)] (2), hedta3− = N-hydrpxyethylenediaminetriacetate, is less discriminating for olefin structures, forming η2 complexes with all eleven olefins and dienes mentioned for studies with 1. However, 2 does not undergo displacement of a carboxylate donor by the second olefin unit of a diene [RuII(hedta)(diene)] complexes possess a pendant non-coordinated olefin and on η2-bound olefin in the complex, indicated by a normal RuII(pac)(olefin)RuII/III wave near 0.55 V.  相似文献   

19.
J. O. D. Coleman  J. M. Palmer 《BBA》1971,245(2):313-320
The ability of triethyltin to inhibit oxidative phosphorylation and electron transport in tightly coupled rat liver mitochondria is very dependent on the pH and the ionic constitution of the assay medium.

1. 1. In an assay medium containing Cl at an alkaline pH, above 7.1, triethyltin inhibited both the ADP stimulated rate of oxygen uptake and the dinitrophenol-induced ATPase (EC 3.6.1.3) but had no effect on the dinitrophenol-stimulated rate of oxygen uptake. If the pH was reduced to below 6.9 the pattern of inhibition changed and both the ADP and dinitrophenol-stimulated rates of oxygen uptake were inhibited by triethyltin.

2. 2. In the absence of Cl in the medium triethyltin inhibited both the ADP-stimulated rate of oxygen uptake and dinitrophenol-induced ATPase and had no effect on the dinitrophenol-stimulated rate of oxygen uptake at either pH 7.4 or 6.6.

3. 3. In either the presence or absence of Cl the ability of triethyltin to inhibit ATP synthesis appears to markedly decrease as the pH is lowered from 7.4 to 6.6.

4. 4. The significance of these observations is discussed in relation to the operation of a Cl/OH antiport in the coupling membrane.

Abbreviations: TMPD, N,N,N′,N′-tetramethylphenylenediamine; FCCP, p-trifluoromethoxyphenylhydrazone  相似文献   


20.

1. 1. Cyanide inhibits the catalytic activity of cytochrome aa3 in both polarographic and spectrophotometric assay systems with an apparent velocity constant of 4·103 M−1·s−1 and a Ki that varies from 0.1 to 1.0 μM at 22 °C, pH 7·3.

2. 2. When cyanide is added to the ascorbate-cytochrome c-cytochromeaa3−O2 system a biphasic reduction of cytochrome c occurs corresponding to an initial Ki of 0.8 μM and a final Ki of about 0.1 μM for the cytochrome aa3−cyanide reaction.

3. 3. The inhibited species (a2+a33+HCN) is formed when a2+a33+ reacts with HCN, when a2+a32+HCN reacts with oxygen, or when a3+a33+HCN (cyano-cytochrome aa3) is reduced. Cyanide dissociates from a2+a33+HCN at a rate of 2·10−3 s−1 at 22 °C, pH 7.3.

4. 4. The results are interpreted in terms of a scheme in which one mole of cyanide binds more tightly and more rapidly to a2+a33+ than to a3+a33+.

Abbreviations: TMPD, N,N,N′,N′-tetramethyl-p-phenylenediamine  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