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.
《Inorganica chimica acta》1987,128(2):169-173
The axial adduct formation of the iron(II) complex of 2,3,9,10-tetraphenyl-l,4,8,11-tetraaza-1,3,8,10-cyclotetradecatetraene (L) with imidazole in dimethyl sulfoxide has been investigated spectrophotometrically at various temperatures and pressures. In the presence of a large excess of imidazole the reaction with the two phases has been observed. The first faster reaction is the formation of the monoimidazole complex of FeL2+, and the second slower reaction corresponds to the formation of the bisimidazole complex. Activation parameters are as follows: for the first step with k1 (25.0°C) = (6.8 ±0.2)×105 mol−1 kg s−1, ΔH31 = 47.5 ± 4.9 kJ mol−1, ΔS31 = 26±16 J K−1 mol−1, and ΔV31 (30.0°C) = 27.2±1.5 cm3 mol−1; for the second step with k2 (25.0°C) = 26.8±0.8 mol−1 kg s−1, ΔH32 = 91.6± 0.8 kJ mol−1, ΔS32 = 90±3 J K−1 mol−1, and ΔV32 (35.0°C) = 21.8±0.9 cm3 mol−1. The large positive activation volumes strongly indicate a dissociative character of the activation process.  相似文献   

2.
The enzyme 2-deoxy-d-ribose-5-phosphate aldolase (DERA) is a useful tool for synthesizing statin side-chain intermediates. In this work, we identified the DERA from Streptococcus suis (SsDERA) by structural and sequence alignment and highly expressed it in Escherichia coli BL21. The recombinant SsDERA had a specific activity of 18.2 U mg−1, KM of 0.8 mM, and Vmax of 32.9 μmol min−1 mg−1 toward 2-deoxy-d-ribose-5-phosphate under the optimal conditions: 40 °C and pH 7.0. The enzyme retained 23.3 % activity after incubation in 200 mM acetaldehyde for 2 h and 58.2 % activity in 100 mM chloroacetaldehyde for 2 h. The enzyme showed moderate activity and aldehyde tolerance compared with reported DERAs. The SsDERA-catalyzed reaction between 200 mM acetaldehyde and 100 mM chloroacetaldehyde generated (3R,5S)-6-chloro-2,4,6-trideoxyhexapyranose in 76 % yield in 8 h. This work provides a new DERA for the synthesis of (3R,5S)-6-chloro-2,4,6-trideoxyhexapyranose, which is a potential candidate for the industrial synthesis of statin intermediates.  相似文献   

3.
The kinetics of the formation of the thiomolybdate ions MoOS32− and MoS42− were determined spectroscopically from the addition of excess sulphide to MoO2S22− in pH buffered media (6–8) at 30 °C. The reverse (hydrolysis) reactions of MoO2S22− and MoOS32− were measured under the same conditions. The reaction rates measured are shown below:
Values of the rate-constants (s−1) obtained at pH 7.0 were k10 2.4 × 10−3, k21 1.5 × 10−5, k30 2.1 × 10−5, k23 6.0 × 10−4, and k34 1.9 × 10−5; where the results are comparable they are in good agreement with those obtained by earlier workers, although different conditions were used. However, in this work it was found that certain reactions had to be mathematically treated as two consecutively occurring reactions. There is also a difference in interpretation of the mechanism of the hydrolysis reactions of the tri- and tetrathio ions. In general the lability towards further S replacement of O atoms, and the reverse reaction, decreased with increased S substitution. All reaction rates increased with increasing H+ ion concentration, mostly this was a linear relationship over the limited pH range examined. The effect of the H+ ion is interpreted in terms of protonation of the oxythiomolybdate ions at an O atom leading to increased lability.  相似文献   

4.
The reactions of PtCl2en or cis-Pt(NH3)2Cl2 and their aqua species with adenine and adenosine were studied by means of ion-pair HPLC. From the chromatograms, it was found that the first binding site of Pt(II) was the N(7) site of adenine under both acidic and neutral conditions. The rates of Pt(II) binding at the (N7) site of adenosine and deoxyadenosine were measured. The rate constants, k1, were obtained for the reactions of PtCl2en or cis-Pt(NH3)2Cl2 with adenosine and deoxyadenosine at pH 3 and 7 over the temperature range 9–25 °C. The k1 values were 6.8–7.7 × 10−4 dm3 mol−1 s−1 at 25 °C. For the aqua species, the rate of [cis-Pt(NH3)2ClH2O]+ with adenosine N(7) was measured. The rate constants, k2 which were found to be smaller than those of hydrolysis, kh, were calculated at pH 3 over the temperature range 25–40 °C. The k2 value obtained at 25 °C was 1.1 × 10−2 dm3 mol−1 s−1, 15 time larger than k1. The activation parameters were also calculated.  相似文献   

5.
Pulse radiolytic studies of α-tocopherol (αTH) oxidation-reduction processes were carried out with low doses (5 Gy) of high-energy electrons in O2−, N2−, and air-saturated ethanolic solutions. Depending on the concentration of oxygen in solution, two different radicals, A· and B·, were observed. The first, A·, was obtained under N2 and results from aTH reaction with solvated electron (kaTH+csolv = 3.4 × 108 mol−1 liter s−1) and with H3C-ĊH-OH, (R·) (kaTH + R· = 5 × 105 mol−1 liter s−1). B·, observed under O2, is produced by αTH reaction with RO2 peroxyl radicals (kaTH + RO2. = 9.5 × 104 mol−1 liter s−1).  相似文献   

6.
《BBA》1986,849(1):121-130
The binding of 3′-O-(1-naphthoyl)adenosinetriphosphate (1-naphthoyl-ATP), ATP and ADP to TF1 and to the isolated α and β subunits was investigated by measuring changes of intrinsic protein fluorescence and of fluorescence anisotropy of 1-naphthoyl-ATP upon binding. The following results were obtained. (1) The isolated α and β subunits bind 1 mol 1-naphthoyl-ATP with a dissociation constant (KD(1-naphthoyl-ATP)) of 4.6 μM and 1.9 μM, respectively. (2) The KD(ATP) for α and β subunits is 8 μM and 11 μM, respectively. (3) The KD(ADP) for α and β subunits is 38 μM μM and 7 μM, respectively. (4) TF1 binds 2 mol 1-naphthoyl-ATP per mol enzyme with KD = 170 nM. (5) The rate constant for 1-naphthoyl-ATP binding to α and β subunit is more than 5 · 104 M−1s−1. (6) The rate constant for 1-naphthoyl-ATP binding to TF1 is 6.6 · 103 M−1 · s−1 (monophasic reaction); the rate constant for its dissociation in the presence of ATP is biphasic with a fast first phase (kA−1 = 3 · 10−3s−1) and a slower second phase (kA−2 < 0.2 · 10−3s−1). From the appearance of a second peak in the fluorescence emission spectrum of 1-naphthoyl-ATP upon binding it is concluded that the binding sites in TF1 are located in an environment more hydrophobic than the binding sites on isolated α and β subunits. The differences in kinetic and thermodynamic parameters for ligand binding to isolated versus integrated α and β subunits, respectively, are explained by interactions between these subunits in the enzyme complex.  相似文献   

7.
《Inorganica chimica acta》1986,121(2):175-183
Chloride anation of trans-Pt(CN)4ClOH2 has been studied with and without Pt(CN)42− present at 25.0°C by use of stopped-flow and conventional spectrophotometry and a 1.00 M perchlorate medium. The rate law in the absence of Pt(CN)42− is Rate=(p1 + p2 [H+] ) [Cl]2 [complex]/(1 + q [Cl]) with p1=(3.0 ± 0.1) × 10−5 M−2s−1, p2=(3.6 ± 0.1) × 10−5 M−3 s−1 and q=(0.62 ± 0.02) M−1. It is compatible with a chloride assistance via an intermediate of the type Cl-Cl-Pt(CN)4···OH22−, in which the reactivity of the aqua ligand is enhanced due to a partial reduction of the platinum. This mechanism of halide assistance is in principle the same as the modified reductive elimination oxidative addition (REOA) mechanism proposed by Poë, in which the intermediate is not split into free halogen, platinum(II) and water, and in which electron transfer not necessarily involves complete reduction to platinum(II). To avoid confusion with complete reductive eliminations, reactions without split of the intermediates are here termed halide-assisted reactions. The pH-dependence indicates acid catalysis via a protonated intermediate ClClPt(CN)4···OH3.The Pt(CN)42−accelerated path has the rate law Rate=
[Cl-] [Pt(CN)42−] [complex] where k=(39.9±0.5) M−2 s−1 and Ka=(4.0±0.2)10−2 M is the protolysis constant of trans-Pt(CN)4ClOH2−.Reaction between PtCl5OH2 and chloride is accelerated by Pt(CN)42− and gives PtCl62− as the reaction product. The rate law is Rate=k [Cl] [Pt(CN)42−] [PtCl5OH2] with k=(5.6 ± 0.2)10−3 M−2 s−1 at 35.0°C and for a 1.50 M perchlorate acid medium. The reaction takes place without central ion exchange. Alternative mechanisms with two consecutive central ion exchanges can be excluded. The role of Pt(CN)42− in this reaction is very similar to that of the assisting halide in the halide assisted anations. [p ]Reaction between trans-Pt(CN)4ClOH2 and PtCl42− gives Pt(CN)42− and PtCl5OH2 as products and has the rate law Rate=k[PtCl42−] [trans-Pt(CN)4ClOH2] with k=(3.32 ± 0.02) M−1 s−1 at 25 °C for a 1.00 M perchloric acid medium. The formation of an aqua complex as the primary reaction product and the rate independent of [Cl] shows that formation of a bridged intermediate of the type Pt(II)Cl4ClPt(IV)(CN)4OH23− is formed in the initial reaction step, not five-coordinated PtCl53−.  相似文献   

8.
The enzymatic production of α-dehydrobiotin (α-DHB), an antibiotic, from biotinyl-CoA using acyl-CoA oxidase and from biotin using a coupling system of biotinyl-CoA synthetase and acyl-CoA oxidase was developed. Acyl-CoA oxidase was found to show activity for biotinyl-CoA. Km and Vmax values of acyl-CoA oxidase for biotinyl-CoA were 75 μM and 3.92 μmol min−1 mg−1, respectively. Optimum reaction conditions for the α-DHB production from biotin were examined. The maximum production of α-DHB (4.29 μmol ml−1) was obtained, when the reaction was carried out at 30°C for 36 h in a mixture consisting of 100 mM potassium phosphate buffer (pH 8.0), 20 mM biotin, 20 mM ATP, 60 mM CoA, 20 mM MgCl2, 2 units of biotinyl-CoA synthetase, 90 units of acyl-CoA oxidase and 25 units of catalase in a total volume of 0.6 ml under aerobic conditions. The product was purified from 14 ml of the reaction mixture and 10 mg of crystals with white needle form were obtained. From NMR, mass spectra and other physical analyses, this compound was identified as (+)-trans-α-DHB.  相似文献   

9.
Complex formation between Pd(II), Pt(II) and iodide has been studied at 25 °C for an aqueous 1.00 M perchloric acid medium. Measurements of the solubility of PdI2(s) in aqueous mercury(II) perchlorate and of AgI(s) and PdI2(s) in aqueous solutions of Pd2+(aq) and Ag+(aq) gave the solubility product of PdI2(s) as Kso=(7±3) × 10−32 M3, which is much smaller than previous literature values.The stability constants β1=[MI(H2O)3+]/([M(H2O)42+][I]) for the two systems were obtained as the ratio between rate constants for the forward and reverse reactions of (i).
The following values of k1 (s−1 M−1), k−1 (s−1) and β1 (M−1) were obtained at 25 °C: (1.14±0.11) × 106, (0.92±0.18), (12±4) × 105 for MPd, and (7.7±0.4), (8.0±0.7) × 10−5, (9.6±1.3) × 104 for MPt. Combination with previous literature data gives the following values of log(β1 (M−1)) to log(β4 (M−4)): 6.08, ∼22, 25.8 and 28.3 for MPd, and 4.98, ∼25, ∼28, and ∼30 for MPt. The present results show that the large overall stability constants β4 observed for the M2+I systems are most likely due to a very large stability of the second complex MI2(H2O)2, which is probably a cis-isomer. A distinct plateau in the formation curve for mean ligand number 2 is obtained both for MPd and Pt. The other iodo complexes are not especially stable compared to those of chloride and bromide.ΔH (kJ mol−1) and ΔS (JK−1 mol−1) for the forward reaction of (i), MPd, are (17.3±1.7) and (−71±5), and for the reverse reaction of (i) MPd, (45±3) and (−95±6), respectively. The kinetics are compatible with associative activation (Ia). The contribution from bond-breaking in the formation of the transition state seems to be less important for Pd than for Pt.  相似文献   

10.
trans-[Ru(NH3)4P(OR)3(H2O)]2+ (R = Me, Pr, iPr, and Bu) reacts with isonicotinamide at second-order- specific rates k1 of 1.2, 2.3, 7.4 and 8.1 M−1 s(25 °C, μ = 0.10 NaCF3COO/CH3COOH), respectively, for R = Me, Pr, iPr and Bu. The products trans- [Ru(NH3)4P(OR)3isn](PF6)2 have been isolated and characterized by micro analysis, cyclic voltammetry, and electronic spectral data. The aquation rates k−1 for the isonicotinamide (isn) derivatives are 5.2 × 10−2, 5.9 × 10−2, 2.0 × 10−1 and 3.4 × 10−1 s−1 for R= Me, Pf, Bu and iPr, respectively. The activation parameters for the forward and backward reactions indicate the same mechanism for all of them. The substitution proceeds by a dissociative mechanism with a significant outer-sphere association of trans-[Ru(NH3)4P(OR)3(H2O)]2+ complexes with isn. Assuming k1 as indicative of the lability of the coordinated water molecule on the monophosphite complexes, the following sequence of increasing trans-effect mav be proposed: P(OMe)3 <P(OEt)3 <P(OPr)3 <P(OiPr)3 <P(OBu)3. The affinity of the monophosphite complexes for isn increases according to P(OMe)3 ⋍ P(OiPr)3 < P(OEt)3 < P(OPr)3 ⋍ P(OBu)3.  相似文献   

11.
《Inorganica chimica acta》1986,121(2):167-174
The reaction of 2,3-tri with CrCl3·6H2O1, dehydrated in boiling DMF, results in the formation of mer-CrCl3(2,3-tri) and anation of hydrolysed solutions of mer-MCl3(2,3,-tri) (M=Co, Cr) with 6 M HCl containing HClO4, forms trans-dichloro- mer-[MCl2(2,3-tri)(OH2)]ClO4·H2O (M=Cr, Co; I, II). trans-Dinitro-mer-[Co(NO2)2(NH3)(2,3-tri)] ClO4 crystallises from the reaction between mer-Co(NO2)3(2,3-tri) and aqueous 7 M ammonia, on addition of NaClO4·H2O, and trans-dichloro-mer-[CoCl2(NH3)(2,3-tri)]ClO4 (III) can be isolated by treatment of the dinitro with 12 M HCl. Reaction of mer-CoCl3(2,3-tri) with C2O42, followed by addition of aqueous NH3 and NaClO4·H2O results in the isolation of racemic mer-[Co(ox)(NH3)(2,3-tri)]ClO4· H2O. This complex was resolved into its enantiomeric forms and treatment of these with SOCl2/MeOH/ HClO4 gave the chiral forms of trans-dichloro-mer- [CoCl2(NH3)(2,3-tri)]ClO4 (R or S at the see-NH center). The rates of loss of the first chloro ligand from these dichloro complexes have been measured spectrophotometrically in 0.1 M HNO3 over a 15 K temperature range to give the following kinetic parameters; (I) kH(298)=7.25 × 10−5 s−1, Ea=78.5 kJ mol−1, δS298#=69 J K−1 mol−1; (II) kH(298)=4.00 × 10−3 s−1, Ea=89.9, δS298#= +87.5; (III) kH(298)=3.09 × 10−4 s−1, Ea=103, δS298#=+27. Treatment of the dichloro cations with Hg2+/HNO3 results in the generation of mer- M(2,3-tri)(OH2)33+ (M=Cr, Co; IV, V) and trans- diaqua-mer-Co(NH3)(2,3-tri)(OH2)23+ (VI). The Co(III) cations isomerise to the fac configuration with (V) Kisom(298) μ=1.0 M)=2.97 × 10−5 s−1, Ea=115, δS298#=+46. (VI) Kisom(298) (μ=1.0 M)=4.13 × 10−5 s−1, Ea=113, δS298#=+52.  相似文献   

12.
A variety of agents cause DNA base alkylation damage, including the known hepatocarcinogen aflatoxin B1 (AFB1) and chemotherapeutic drugs derived from nitrogen mustard (NM). The N7 site of guanine is the primary site of alkylation, with some N7-deoxyguanosine adducts undergoing imidazole ring-opening to stable mutagenic N5-alkyl formamidopyrimidine (Fapy-dG) adducts. These adducts exist as a mixture of canonical β- and unnatural α-anomeric forms. The β species are predominant in double-stranded (ds) DNA. Recently, we have demonstrated that the DNA glycosylase NEIL1 can initiate repair of AFB1-Fapy-dG adducts both in vitro and in vivo, with Neil1−/− mice showing an increased susceptibility to AFB1-induced hepatocellular carcinoma.Here, we hypothesized that NEIL1 could excise NM-Fapy-dG and that NEIL3, a closely related DNA glycosylase, could excise both NM-Fapy-dG and AFB1-Fapy-dG. Product formation from the reaction of human NEIL1 with ds oligodeoxynucleotides containing a unique NM-Fapy-dG followed a bi-component exponential function under single turnover conditions. Thus, two adduct conformations were differentially recognized by hNEIL1. The excision rate of the major form (∼13.0 min−1), presumed to be the β-anomer, was significantly higher than that previously reported for 5-hydroxycytosine, 5-hydroxyuracil, thymine glycol (Tg), and AFB1-Fapy-dG. Product generation from the minor form was much slower (∼0.4 min−1), likely reflecting the rate of conversion of the α anomer into the β anomer. Mus musculus NEIL3 (MmuNEIL3Δ324) excised NM-Fapy-dG from single-stranded (ss) DNA (turnover rate of ∼0.4 min−1), but not from ds DNA. Product formation from ss substrate was incomplete, presumably because of a substantial presence of the α anomer. MmuNEIL3Δ324 could not initiate repair of AFB1-Fapy-dG in either ds or ss DNA. Overall, the data suggest that both NEIL1 and NEIL3 may protect cells against cytotoxic and mutagenic effects of NM-Fapy-dG, but NEIL1 may have a unique role in initiation of base excision repair of AFB1-Fapy-dG.  相似文献   

13.
《BBA》1986,851(3):361-368
Absorbance changes in the picosecond region were studied in isolated reaction centers of the green photosynthetic bacterium Chloroflexus aurantiacus upon selective excitation of the primary electron donor, P, at 870 nm. The results indicate that the first observed state is an excited state of P (P1) which appears to transfer an electron to a bacteriochlorophyll a molecule absorbing at 812 nm (B1) in 10 ± 2 ps as indicated by a bleaching at this wavelength. This reaction is followed by a rapid electron transfer (3 ± 1 ps) from B1 to bacteriopheophytin a, so that the fraction of reaction centers in the state P+B1 remains small during the experiment. An apparent bleaching at 925 nm is ascribed to stimulated emission from excited P, which emission disappears upon formation of P+. The difference between these time constants for electron transfer and those observed for the same reactions in reaction centers of the purple photosynthetic bacterium Rhodopseudomonas (Rhodobacter) sphaeroides is discussed in terms of the energy difference between P1 and P+B1, which appears to be larger for C. aurantiacus.  相似文献   

14.
Red blood cell (rbc) carbon dioxide transport was examined in vitro in three teleosts (Oncorhynchus mykiss, Anguilla anguilla, Scophthalmus maximus) and an elasmobranch (Scyliorhinus canicula) using a radioisotopic assay that measures the net conversion of plasma HCO3 to CO2. The experiments were designed to compare the intrinsic rates of rbc CO2 excretion and the impact of haemoglobin oxygenation/deoxygenation among the species.Under conditions simulating in vivo levels of plasma HCO3 and natural haematocrits, the rate of whole blood CO2 excretion varied between 14.0 μmol ml−1 h−1 (S. canicula) and 17.6 μmol ml−1 h−1 (O. mykiss). The rate of CO2 excretion in separated plasma was significantly greater in the dogfish, S. canicula. The contribution of the rbc to overall whole blood CO2 excretion was low in the dogfish (46 ± 6%) compared to the teleosts (trout, 71 ± 4%; turbot, 64 ± 5%; eel, 55 ± 3%).To eliminate the naturally occurring differences in haematocrit and plasma [HCO3] as inter-specific variables, the rates of whole blood CO2 excretion were determined in blood that had been resuspended to constant [HCO3] (5 mmol−1) and haematocrit (20%) in appropriate teleost and elasmobranch Ringer solutions. Under such normalized conditions, the rate of whole blood CO2 excretion was significantly higher in the turbot (22.4 ± 1.3 μmol ml−1 h−1) in comparison to the other species (16.4–18.4 μmol ml−1 h−1) and thus revealed a greater intrinsic rate of rbc CO2 excretion in the turbot.To study the contribution of Bohr protons, the rates of whole blood CO2 excretion were assessed in blood subjected to rapid oxygenation during the initial phase of the 3 min assay period. Rapid oxygenation significantly enhanced the rate of CO2 excretion in the teleosts but not in the elasmobranch. The extent of the increase provided by the rapid oxygenation of haemoglobin was a linear function of the extent of the Haldane effect, as quantified in each species from in vitro CO2 dissociation (combining) curves. Under steady-state conditions, deoxygenated blood exhibited greater rates of CO2 excretion than oxygenated blood in the teleosts but not in the elasmobranch. As a consequence of the Haldane effect, rbc intracellular pH was increased in the teleosts by deoxygenation but was unaltered in the elasmobranch.The results, by extrapolation, suggest that the rates of CO2 excretion in vivo are influenced by the magnitude of the Haldane effect and the extent of haemoglobin oxygenation during gill transit in addition to the intrinsic rate at which the rbc converts plasma HCO3 to CO2.  相似文献   

15.
A kinetic study of the oxidation of (hydroxyethyl)ferrocene (HEF) by [2-pyridylmethylbis(2-ethyl-thioethyl)ainine]copper(II) (Cu(pmas)2+) is reported, with the objective of documenting the influence of the two thioether sulfur ligands on the electron transfer rate. Both reactants exhibit a first-order dependence at pH 6, I = 0.1 M(NaNO3); k(25°C) = 1.3 × 104M−1sec−1, ΔH3 = 10.1 kcal/mole, ΔS3 = −6 eu. The apparent Cu(pmas)2+/+ self-exchange electron transfer rate constant calculated from this reaction on the basis of relative Marcus theory (4.7 × 101M−1 sec−1) agrees well with previous findings on ferrocytochrome c, Fe(CN)64−, and Ru(NH3)5py2+ oxidations. Spectrophotometric titrations of Cu(pmas)2+ and Cu(tmpa)2+ (tmpa = tris(2-pyridylmethyl)amine) with azide ion showed that both Cu(pmas)N3)+ (Kf1 = 3.1 × 103M−1) and Cu(pmas)(N3)2 (Kf2 = 3.5 × 101M−1) but Cu(tmpa)(N3)+ (Kf = 6.6 × 102M−1) are formed up to 0.15 M N3 (25°C, pH 6, I = 0.2 M), suggesting that a thioether sulfur atom is displaced in the uptake of a second N3 ion by Cu(pmas)(N3)+. The effect of thioether sulfur displacement by azide ion on the HEF-Cu(pmas)2+ reaction rate may be understood entirely through the tendency of N3 to shift the position of the redox equilibrium towards the reactant side, without invoking any special role for the sulfur ligand in promoting electron transfer reactivity.  相似文献   

16.
《Inorganica chimica acta》1986,115(2):223-227
The exchange reaction of acac(acetylacetonate) in UO2(acac)2dmf (dmf=N,N-dimethylformamide) in o-dichlorobenzene has been studied by the NMR line-broadening method. The exchange rate depends on the concentration of the enol isomer of acetylacetone in its low region, and approaches the limiting value in its high region. It is proposed that the exchange reaction proceeds through the mechanism in which the dissociation of one end of the chelated acac is the rate-determining step. The kinetic parameters for this step are as follows: k (25 °C)=5.03 × 10−3 s−1, ΔH3=91.6 ± 3.8 kJ mol−1, and ΔS3 =17.2 ± 10.5 JK−1 mol−1. The exchange rate becomes slower by the addition of free DMF. This may be due to the competition of DMF with the enol isomer of acetylacetone in attacking a four-coordinated intermediate in the equatorial plane.  相似文献   

17.
《BBA》1985,807(1):81-95
(1) The apparent Km for nitrate of the electron-transport system in intact cells of Paracoccus denitrificans was less than 5 μM. In contrast the apparent Km for nitrate of inverted membrane vesicles oxidising NADH was greater than 50 μM. When azide, a competitive inhibitor, was present, the apparent Km observed with the vesicles was raised to 0.64 mM, consistent with values previously reported for purified preparations of the reductase. In membrane vesicles the nitrate reductase is probably not rate-limiting for NADH-nitrate oxido-reductase activity, and thus a lower limit for Km (NO3) is obtained. It is suggested that the very low Km (NO3) in intact cells must arise from either a transport process or a nitrate-specific pore that allows access of nitrate directly to the active site of its reductase from the periplasm. (2) The swelling of spheroplasts has been studied under both aerobic and anaerobic conditions to probe possible mechanisms of nitrate and nitrite transport across the plasma membrane of P. denitrificans. Nitrate reductase was inhibited by azide to prevent reduction of internal nitrate. No evidence for operation of either nitrate-nitrite antiport or proton-nitrate symport was obtained. (3) Measurements from the fluorescence intensity of 8-anilino-naphthalene-1-sulphonate of the rates of decay of diffusion potentials generated by addition of potassium salts to valinomycin-treated plasma membrane vesicles from P. denitrificans showed that the permeability of the membrane to anions is SCN > NO3, NO2, pyruvate, acetate > CI > SO42−. In the presence of a protonophore the rate of decay of the diffusion potential was considerably enhanced with potassium acetate or potassium nitrite, but not with potassium salts of nitrate, chloride or pyruvate. This result indicates that HNO2 and CH3COOH can rapidly and passively diffuse across the cell membrane. This finding suggests that transport systems for nitrite are in general probably not required in bacteria. The failure of a protonophore to enhance the dissipation of the diffusion potential generated by potassium nitrate is evidence against the operation of a proton-nitrate symporter. (4) Low concentrations of added nitrite very strongly inhibit electron flow to oxygen in anaerobically grown cells, provided that they have been treated with Triton X-100 or an uncoupler. This inhibition is not observed with aerobically grown cells. It is concluded that the inhibitory species is a reaction product or an intermediate of the nitrite reductase reaction. The requirement for collapse of protonomotive force by uncoupler or permeabilising the plasma membrane suggests that any such species could be negatively charged. Nitroxyl anion (NO) can be considered, as its conjugate acid is a postulated intermediate between nitrite and nitrous oxide; nitroxyl anion can bind to heme centres to give nitrosyl derivatives. (5) The basis for the ability of permeabilised, but not intact, cells of P. denitrificans to reduce oxygen and nitrate simultaneously is discussed.  相似文献   

18.
《Inorganica chimica acta》1988,150(1):81-100
The (NH3)5CoOC(NH2)23+ ion is consumed in water according to the rate law k(obs.) = k1 + k2[OH], where k1 = 4.0 × 10−5 s−1 and k2 = 14.2 M−1 s−1 (0–0.1 M [OH];μ = 1.1 M, NaClO4, 25 °C). A hitherto unrecognized intramolecular O- to N- linkage isomerization reaction has been detected. In strongly acid solution only aquation to (NH3)5CoOH23+ is observed, but in 0.1–1.0 M [OH], 7% of the directly formed products is the urea-N complex (NH3)5CoNHCONH22+ which has been isolated. In the neutral pH region a much greater proportion (25%) of the products is the urea-N species. These results are interpreted in terms of an urea-O to urea-N linkage isomerization reaction competing with hydrolysis for both spontaneous (k1) and base-catalyzed (k2) pathways; the rearrangement is not observed in strongly acidic solution (pH ⩽ 1) because the protonated N-bonded isomer (pKa ≈ 3) is unstable with respect to the O-bonded form. The appearance of the isomerization pathway as the pH is raised in the 0–6 region is commensurate with a rate increase which cannot be attributed to a contribution from the base catalysis term k2[OH]. It is argued that this observation establishes, for the spontaneous pathway, that hydrolysis and linkage isomerization are separate reaction pathways — there is no common intermediate. The product distribution and rate data lead to the complete rate law, k(obs.) = k1 + k2[OH] = (ks + kON) + (kOH + kON) [OH] for the reactions of the O-bonded isomers, where ks, kOH are the specific rates for hydrolysis, and kON, kON are the specific rates for O- to N-linkage isomerization, by spontaneous and base-catalyzed pathways respectively; kON = 1.3 × 10−5 s−1 and kON = 1.1 M−1 s−1 (μ = 1.0 M, NaClO4, 25 °C). The O- to N- linkage isomerization has been observed also for complexes of N-methylurea, N,N-dimethylurea and N-phenylurea, but not for the N,N′-dimethylurea species. There is an approximately statistical relationship among the data for −NH2 capture (versus H2O), while −NHR and −NR2 do not compete with water as nucleophiles for Co(III) in either the spontaneous or base-catalyzed hydrolysis processes. For each urea-O complex, O- to N-isomerization is a more significant parallel reaction in the spontaneous as opposed to the base-catalyzed pathway. This is interpreted as being indicative of more associative character in the spontaneous route to products, a conclusion supported by other evidence. Some activation parameter data have been recorded and the effect of the N-substitution on the rates of solvolysis (H2O, Me2SO) is discussed. The urea-N complexes have been isolated as their deprotonated forms, [(NH3)5CoNHCONRR′](ClO4)2·xH2O (R,R′ = H, CH3). They are kinetically inert in neutral to basic solution but in acid they protonate (H2O, pKa 2–3; μ = 1.0 M, 25 °C) and then isomerize rapidly back to their O-bonded forms. Some solvolysis accompanies this N- to O-rearrangement in H2O and Me2SO. Specific rates and activation parameters are reported. The kinetic data follow a rate law of the form kNO(obs.) = (k + kNO)[H+]/(Ka + [H+]) and the active species in the reaction is the protonated form; k, kNO are the specific rates for hydrolysis and isomerization, respectively. Proton NMR data establish that the site of protonation (in Me2SO) is the cobalt-bound nitrogen atom. For the unsubstituted urea species (NH3)5CoNH2CONH23+, diastereotopic exo-NH2 protons arising from restricted rotation about the CN bond are observed. The relevance to the mechanism of the linkage isomerization process is considered. 13C and 1H NMR and electronic absorption spectral data are presented, and distinctions between linkage isomers and the solution structures (electronic and conformational) are discussed. The urea-N/urea-O complex equilibrium is governed by the relation KNO(obs.) = KNO[H+]/[H+](Ka), where KNO is the equilibrium constant = [(NH35Co(urea-O)3+]/[(NH3)5Co(urea-N)3+]. Values for KNO(=kNO/kON = 260 and pKa ≈ 3 for the NH2CONH2 system are consistent with the stability of the N-isomer in feebly acidic to basic solution (e.g. pH 6, KNO(obs.) = 2.6 × 10−2) and instability in acid solution (e.g. pH 1, KNO(obs.) = 240). The equilibrium data for this and other urea complexes of (NH3)5Co(III) are contrasted with the result for the analogous Rh(III)NH2CONH2 system KNO ≈ 1).  相似文献   

19.
《Inorganica chimica acta》1987,130(2):183-184
cis,cis,trans-[PtIV(NH3)2Cl2(OH)2] reacts reversibly with ascorbic acid to give dehydroascorbic acid and mainly cis-[PtII(NH2Pri)2Cl2]. The parameters for the forward reaction are: kf = 0.584 M s at 37.0 °C, ΔHf = 108.6 −+ 6.4 kJ mol−1 andΔSf = 101 −+ 22 J K−1 mol−1.  相似文献   

20.
Control of denitrification enzyme activity in a streamside soil   总被引:3,自引:0,他引:3  
Progress curve analysis of NO3 and NO2 reduction in surface soil samples from a streamside soil gave Km values of 4.24 and 6.33 μM, and Vmax values of 2.16 and 1.83 μmol l−1 min−1, respectively. Recoveries of reduced NO3 and NO2 as gaseous N averaged 82 and 108%. The unrecovered NO3-N was presumably dissimilated to NH4+-N. The idenitification enzyme activity (DEA) was examined throughout a year and showed seasonal and spatial variabilities of only 10% to 26%, suggesting a high persistency of denitrifying enzymes. Soil moisture and DEA correlated significantly (r = 0.7671; P<0.01). The DEA in saturated subsoil also showed a relatively little variation, with spatial variabilities of between 28 and 38%. Amendment with NO3 rarely enhanced the acctivity more than two-fold at either depth. Addition of glucose increased the activity 2.3 and 2.5 times in the surface soil and suboil respectively, indicating a moderate carbon limitation of denitrification. The activation energy of DEA was found to be 64.9 kJ mol−1 and Q10 values for the 2–12°C and 12–22°C temperature ranges were 2.71 and 2.53, respectively. Extrapolation suggested there would be a 4.4-fold increase in DEA if the temperature was changed from 0 to 15°C. Substrate diffusion limited the denitrification 10 to 25 fold.Thus, under anaerobic moist conditions it appears that changes in denitrification might primarily be due to varying diffusion of substrates into the anaerobic soil centers. Over a year, fluctuations in DEA, temperature changes and fluctuations of electron-acceptor and -donor supply will only have a minor effect on natural denitrification activity.  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