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 kinetics and mechanism of a linear trihydroxamic acid siderophore (deferriferrioxamine B, H4DFB+) ligand exchange with Al(H2O)63+ to form mono(deferriferrioxamine B)aluminum(III) (Al(H2O)4H3DFB)3+ have been investigated at 25 °C over the [H+] range 0.001−1.0 M and I = 2.0 M (HClO4/NaClO4) by 27Al NMR. Kinetic results are consistent with Al(H2O)4(H3DFB)3+ formation and dissociation proceeding through a parallel path mechanistic scheme involving Al(H2O)63+(k2/k−1) and Al(H2O)5(OH)2+(k2/k−2) where k1 = 0.13 M−1 s−1, k−1 = 8.7 × 10−3 M−1 s−1, k2 = 2.7 × 103 M−1 s−1, and k−2 = 9.6 × 10−4 s−1. Relative complex formation rates at Al(H2O)63+ and Al(H2O)5OH2+, and comparison with kinetic data for a series of synthetic hydroxamic acids, suggest that an interchange mechanism is operative. These results are also discussed in relation to kinetic data for the corresponding iron(III)-deferriferrioxamine B system.  相似文献   

2.
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   

3.
The reactions of copper(II)-ahphatic polyamine complexes with cysteine, cysteine methyl ester, penicillamine. and glutathione have been investigated, with the goal of understanding the relationship between RS?-Cu(II) adduct structure and preferred redox decay pathway. Considerable mechanistic flexibility exists within this class of mercapto ammo acid oxidations, as changes in the rate law could be induced by modest variations in reductant concentration (at fixed [Cu(II)]o), pH, and the structure of the redox partners. With excess cysteine present at 25°C, pH 5 0, I = 0 2 M (NaOAc), decay of 1:1 cys-S?-Cu(II) transient adducts was found to be first order in both cys-SH and transient. Second-order rate constants characteristic of Cu(dien)2+ (6 1 × 103M?1sec?1), Cu(Me5dien)2+ (2.7 × 103M?1 sec?1), Cu(en)22+ (2.1 × 103M?1 sec?1), and Cu(dien)22+ (4.7 × 103 M?1 sec ?1) are remarkably similar, considering substantial differences in the composition and geometry of the oxidant first coordination sphere. A mechanism involving attack of cysteine on the coordinated sulfur atom of the transient, giving a disulfide anion radical intermediate, is proposed to account for these results Moderate reactivity decreases in the cysteine-Cu(dien)2+, Cu(Me5dien)2+ reactions with increasing [H+] (pH 4–6) reflect partial protonation of the polyamine ligands. A very different rate law, second order in the RS?-Cu(II) transient and approximately zeroth order in mercaptan, applies in the pH 5.0 oxidations of cysteine methyl ester, penicillamine, and glutathione by Cu(dien)2+ and Cu(Me5dien)2+. This behavior suggests the mtermediacy of di-μ-mercapto-bridged binuclear Cu(II) species, in which a concerted two-electron change yields the disulfide and Cu(I) products. Similar hydroxo-bridged intermediates are proposed to account for the transition from first- to second-order transient dependence in cysteine oxidations by Cu(dien)2+ and Cu(Me5dien)2+ as the pH is increased from 5 to 7. Yet another rate law, second order in transient and first order in cysteine, applies in the pH 5.0 oxidation of cysteine by Cu(Me6tren)2+ (k(25°C) 7.5 × 107 M?2 sec?1, I = 0.2 M). Steric rigidity of this trigonal bipyramidal oxidant evidently protects the coordinated sulfur atom from attack in a RSSR?-forming pathway. Formation of a coordinated disulfide in the rate-determining step is purposed, coupled with attack of a noncoordinated cysteine molecule on a vacated coordination position to stabilize the (Me6(tren)Cu(I) product.  相似文献   

4.
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   

5.
《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   

6.
《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   

7.
Porphyrins are a chemical class that is widely used in drug design. Cationic porphyrins may bind to DNA guanine quadruplexes. We report the parameters of the binding of 5,10,15,20-tetrakis(N-carboxymethyl-4-pyridinium) porphyrin (P1) and 5,10,15,20-tetrakis(N-etoxycarbonylmethyl-4-pyridinium) porphyrin (P2) to antiparallel telomeric G-quadruplex formed by d(TTAGGG)4 sequence (TelQ). The binding constants (K i ) and the number of binding sites (N j ) were determined from absorption isotherms generated from the absorption spectra of complexes of P1 and P2 with TelQ. Compound P1 demonstrated a high affinity to TelQ (K i = (40 ± 6) × 106 M?1, N 1 = 1; K 2 = (5.4 ± 0.4) × 106 M?1, N 2 = 2). In contrast, the binding constants of P2-TelQ complexes (K 1 = (3.1 ± 0.2) × 106 M?1, N 1 = 1; K 2 = (1.2 ± 0.2) × 106 M?1, N 2 = 2) were one order of magnitude smaller than the corresponding values for P2-TelQ complexes. Measurements of the quantum yield and fluorescence lifetime of the drug’s TelQ complexes revealed two types of binding sites for P1 and P2 on the quadruplex oligonucleotide. We concluded that strong complexes can result from the interaction of the porphyrins with TTA loops whereas the weaker complexes are formed with G-quartets. The altered TelQ conformation detected by the circular dichroism spectra of P1-TelQ complexes can be explained by the disruption of the G-quartet. We conclude that peripheral carboxy groups contribute to the high affinity of P1 for the antiparallel telomeric G-quadruplex.  相似文献   

8.
Copper(II) complexes of a series of linear pentadentate ligands containing two benzimidazoles, two thioether sulfurs and a amine nitrogen, viz. N,N-bis{4-(2″-benzimidazolyl)(methyl)-3-thiabutyl}amine(L1), N,N-bis{4-(2″-benzimidazolyl)(methyl)-3-thiabutyl}N-methylamine (L2), 2,6-bis{4-(2″-benzimidazolyl)(methyl)-3-thiabutyl}pyridine(L3), N,N-bis{4-(2″-benzimidazolyl)-2-thiabutyl}amine (L4), N,N-bis{4-(2″-benzimidazolyl)-2-thiabutyl}N-methylamine (L5) and 2,6-bis{4-(2″-benzimidazolyl)-2-thiabutyl}-3pyridine (L6) have been isolated and characterized by electronic absorption and EPR spectroscopy and cyclic and differential pulse voltammetry. Of these complexes, [Cu(L1)](BF4)2 (1) and [Cu(L2)](BF4)2 (4) have been structurally characterized by X-ray crystallography. The coordination geometries around copper(II) in 1 and 4 are described as trigonal bipyramidal distorted square based pyramidal geometry (TBDSBP). The distorted CuN3S basal plane in them is comprised of amine nitrogen, one thioether sulphur and two benzimidazole nitrogens and the other thioether sulfur is axially coordinated. The ligand field spectra of all the complexes are consistent with a mostly square-based geometry in solution. The EPR spectra of complexes [Cu(L1)](BF4)2 (1), [Cu(L1)](NO3)2 (2), [Cu(L2)](BF4)2 (4) and [Cu(L3)](ClO4)2 (6) are consistent with two species indicating the dissociation/disproportionation of the complex species in solution. All the complexes exhibit an intense CT band in the range 305-395 nm and show a quasireversible to irreversible CuII/CuI redox process with relatively positive E1/2 values, which are consistent with the presence of two-coordinated thioether groups. The addition of N-methylimidazole (mim) replaces the coordinated thioether ligands in solution, as revealed from the negative shift (222-403 mV) in the CuII/CuI redox potential. The present study reveals that the effect of incorporating an amine nitrogen donor into CuN2S2 complexes is to generate an axial copper(II)-thioether coordination and also to enforce lesser trigonality on the copper(II) coordination geometry.  相似文献   

9.
Two novel copper (II) complexes [Cu(TFP)(Gly)Cl] ⋅ 2H2O complex ( 1 ) and [Cu(TFP)(His)Cl] ⋅ 2H2O complex ( 2 ) are synthesized, where TFP stands for trifluropromazine, Gly. represents glycine, and His. is histidine. Chemical composition, IR, mass spectra, and magnetic susceptibility tests are performed. Complex binding with macromolecules was investigated using UV-vis, viscosity, gel electrophoresis, and fluorescence quenching. Fluorescence spectroscopy revealed that each complex could replace ethidium bromide (EB). These complexes exhibit grooved, non-covalent, and electrostatic interactions with CT-DNA. Spectroscopy analysis of the BSA interaction showed that complexes bind to protein (Kb values for ( 1 ) is 5.89×103 M−1 and for ( 2 ) is 9.08×103 M−1) more strongly than CT-DNA (Kb values for ( 1 ) is 5.43×103 M−1 and for ( 2 ) is 7.17×103 M−1). Molecular docking analysis and spectral absorption measurements showed high agreement. Antimicrobial, antioxidant, and anti-inflammatory properties were tested in vitro. The druggability of complex ( 2 ) should be tested in vivo as it is more biologically active.  相似文献   

10.
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   

11.
Equilibrium unfolding (folding) studies reveal that the autoregulatory RNA pseudoknots derived from the bacteriophage T2 and T4 gene 32 mRNAs exhibit significant stabilization by increasing concentrations of divalent metal ions in solution. In this report, the apparent affinities of exchange inert trivalent Co(NH3) have been determined, relative to divalent Mg2+, for the folded, partially folded (Kf), and fully unfolded (Ku) conformations of these molecules. A general nonspecific, delocalized ion binding model was developed and applied to the analysis of the metal ion concentration dependence of individual two‐state unfolding transitions. Trivalent Co(NH3) was found to associate with the fully folded and partially unfolded pseudoknotted forms of these RNAs with a Kf of 5–8 × 104 M−1 in a background of 0.10 M K+, or 3‐ to 5‐fold larger than the Kf obtained for two model RNA hairpins and hairpin unfolding intermediates, and ≈ 40–50‐fold larger than Kf for Mg2+. The magnitude of Kf was found to be strongly dependent on the monovalent salt concentration in a manner qualitatively consistent with polyelectrolyte theory, with Kf reaching 1.2 × 105 M−1 in 50 mM K+. Two RNA hairpins were found to have affinities for Co(NH3) and Ru(NH3) of 1–2 ×104 M−1, or ≈ 15‐fold larger than the Kf of ∼ 1000 M−1 observed for Mg2+. Additionally, the Ku of 4,800 M−1 for the trivalent ligands is ≈ 8‐fold larger than the Ku of 600 M−1 observed for Mg2+. These findings suggest that the T2 and T4 gene 32 mRNA pseudoknots possess a site(s) for Mg2+ and Co(NH3) binding of significantly higher affinity than a “duplexlike” delocalized ion binding site that is strongly linked to the thermodynamic stability of these molecules. Imino proton perturbation nmr spectroscopy suggests that this site(s) lies near the base of the pseudoknot stem S2, near a patch of high negative electrostatic potential associated with the region where the single loop L1 adenosine crosses the major groove of stem S2. © 1999 John Wiley & Sons, Inc. Biopoly 50: 443–458, 1999  相似文献   

12.
A novel ruthenium(II) complex of dipyridophenazine (DPPZ) with the ancillary ligand imidazole[4,5-f] [1,10]phenanthroline (IP), [Ru(IP)2(DPPZ)] (PF6)2, has been synthesized and characterized by elemental analysis, 1D and 2D 1H NMR, fast-atom bombardment mass spectra (FABMS), electronic spectroscopy and cyclic voltammetry. The DNA-binding properties of the complex were studied by spectroscopic methods. The intrinsic binding constant, K =2.1 × 107M−1, of the complex to calf thymus DNA has been determined by absorption titration in 5 mmol dm−3 Tris-HCl, 50 mmol dm−3 NaCl buffer (pH 7.0). The excited state lifetimes and luminescence quenching with [Fe(CN)6]4− as the quencher in the presence of DNA were also tested and mono-exponentiality was observed for the emission decay curves. Viscosity measurements together with the optical titrations unambiguously proved that the complex bound with DNA intercalatively and that the binding affinity to DNA was several times larger than that of the parent complex [Ru(bpy)2(DPPZ)]2+.  相似文献   

13.
Activated white cells use oxidants generated by the heme enzyme myeloperoxidase to kill invading pathogens. This enzyme utilizes H2O2 and Cl, Br, or SCN to generate the oxidants HOCl, HOBr, and HOSCN, respectively. Whereas controlled production of these species is vital in maintaining good health, their uncontrolled or inappropriate formation (as occurs at sites of inflammation) can cause host tissue damage that has been associated with multiple inflammatory pathologies including cardiovascular diseases and cancer. Previous studies have reported that sulfur-containing species are major targets for HOCl but as the reactions are fast the only physiologically relevant kinetic data available have been extrapolated from data measured at high pH (>10). In this study these values have been determined at pH 7.4 using a newly developed competition kinetic approach that employs a fluorescently tagged methionine derivative as the competitive substrate (k(HOCl + Fmoc-Met), 1.5×108 M−1 s−1). This assay was validated using the known k(HOCl + NADH) value and has allowed revised k values for the reactions of HOCl with Cys, N-acetylcysteine, and glutathione to be determined as 3.6×108, 2.9×107, and 1.24×108 M−1 s−1, respectively. Similar experiments with methionine derivatives yielded k values of 3.4×107 M−1 s−1 for Met and 1.7×108 M−1 s−1 for N-acetylmethionine. The k values determined here for the reaction of HOCl with thiols are up to 10-fold higher than those previously determined and further emphasize the critical importance of reactions of HOCl with thiol targets in biological systems.  相似文献   

14.
  • 1.1. The interaction of haemopexin and albumin with TPPS4 was studied by measuring the absorption and fluorescence spectra. Haemopexin was found to have one strong TPPS4 binding center (Ka = 3 × 107M−1).
  • 2.2. Haem-haemopexin complex appears to have no specific binding site for TPPS4. Occupation of the specific binding center of haemopexin molecule by a haem abolishes TPPS4 binding.
  • 3.3. Albumin was found to possess one strong TPPS4 binding center (Ka = 3 × 106M−1) besides two or three weak binding sites (Ka = 2 × 105M−1).
  • 4.4. Haern-albumin complex possesses only one weak TPPS4 binding site (Ka = 7 × lO5M−1). These observations suggest identity of primary binding sites of TPPS4 and haem on albumin molecule.
  相似文献   

15.
《Inorganica chimica acta》1988,147(2):189-197
Complexes of the M(en)3Ag2(CN)4 (M = Ni, Zn, Cd) and M(en)2Ag2(CN)4 (M = Ni, Cu, Zn, Cd) type were prepared and identified by elemental analysis, infrared spectroscopy, measurement of magnetic susceptibility, and X-ray powder diffractometry. The crystal structures of Ni(en)3Ag2(CN)4 (I) and Zn(en)2Ag2(CN)4 (II) were determined by the method of monocrystal structure analysis. Complex I crystallizes in the space group C2/c, a = 1.2639(5), b = 1.3739(4), c = 1.2494(4) nm, β = 113.25(4)°, Dm = 1.86(1), Dc = 1.86 gcm−3 Z = 4, R = 0.0429. The crystal structure of I consists of complex cations [Ni(en)3]2+ and complex anions [Ag(CN)2]. Complex II crystallizes in the space group I2/m, a = 0.9150(3), b = 1.3308(4), c = 0.6442(2) nm, β = 95.80(3)°, Dm = 2.14(1), Dc = 2.15 gcm−3, Z = 2, R = 0.0334. Its crystal structure consists of infinite, positively charged chains of the [-NCAgCNZn- (en)2]nn+ type and isolated [Ag(CN)2] anions. The atoms of Ag are positioned parallely to the z axis and the AgAg distance is equal to 0.3221(2) nm.  相似文献   

16.
《Inorganica chimica acta》1986,121(2):223-228
One isomer of [CrCl(N-Me-tn)(dien)]ZnCl4 has been isolated from the reaction of CrCl3·6H2O, dehydrated in DMF, with the polyamines N-methyl- 1,3-diaminopropane and diethylenetriamine. This complex is isomorphous with δλ-(R,S)usft-[CoCl(N-Me-tn)(dien)] ZnCl4 and thus has the unsym-fac- configuration with the N-Me group trans to the sec-NH group of the coordinated triamine. The Cr(III) complex has been resolved with NH4BCS and the chiroptical parameters of (-)488-[CrCl(N-Me-tn)(dien)]- ZnCl4, derived from the less soluble diastereoisomeride by metathesis, are similar to those obtained for the less soluble (-)534-λ-(S)-a,cb,edf-Co(III) analogue, of known absolute configuration. Kinetic parameters for the rates of thermal aquation (μ= 1.0 M, HClO4) and Hg2+-assisted chloride release (μ= 1.0 M) for usft-[CrCl(N-Me-tn)(dien)]ZnCl4 are kH= 3.7 × 10−6 s−1, Ea=93 ± 8 kJ mol−1, ΔS2984t#= −45 ± 16 J K−1 mol−1 and kHg=2.01 × 10−3 M−1 s−1, Ea=64.2 ± 3.3, ΔS298#=−89.5 ± 7, respectively at 298 K.  相似文献   

17.
《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   

18.
  • 1.1. Phoronis architecta hemoglobin is composed of four distinct hemoglobin subunits with minimum MW's of 16–17,000 or 17–19,000 daltons. All four hemoglobins are monomeric when oxygenated. Two of the monomers combine to form dimers when bound with carbon monoxide.
  • 2.2. In cellulo, Phoronis architecta hemoglobin has a half-saturation (P50) value of 1.3 ± 0.1 mm Hg, shows cooperative oxygen binding (Hill coefficient = 2.7 ± 0.3), and no Bohr effect from pH 6.6 to 7.9. In vitro, the hemoglobin has a P50 of 0.76 ± 0.21 mm Hg but shows no cooperativity (0.90 ± 0.15 (SD)).
  • 3.3. The oxygen dissociation constant (Koff) from hemoglobin is 2.7 ± 0.2 sec−1, and the computed oxygen association constant (Kon) is 2.5 × 106 M−1 · sec−1 (1.9–3.6 × 106 M−1 · sec−1).
  相似文献   

19.
《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   

20.
《Inorganica chimica acta》1988,141(2):211-220
The reaction of CrCl3 · 6H2O (dehydrated in DMSO) with 1,5,9-triazanonane (3,3-tri) gives mer- CrCl3(3,3-tri), the configuration being established by isomorphism with the corresponding Co(III) complex. This non-electrolyte is hydrolyzed in aqueous acidic solution and mer-[CrCl2(3,3-tri)- (OH2)]ClO4 can be isolated by anation with HCl in the presence of HClO4. Reaction of mer-CrCl3- (3,3-tri) in DMF with diamines produces complexes of the type [CrCl(diamine)(3,3-tri)] Cl2 [diamine= 1,2-diaminoethane (en), 1.2-diaminopropane (pn), 1,3-diaminopropane (tn), 2,2-dimethyl-1,3-diaminopropane (Me2tn) and cyclohexanediamine (chxn, cis plus trans mixture; two isomers A and B)] and these have been characterized as the ZnCl42− salts. The configuration of the triamine ligand in these complexes has been established as mer-(H↓)- by a single crystal X-ray analysis of [CrCl(en)(3,3-tri)]- ZnCl4, monoclinic, P21, a=7.932, b= 14.711, c= 8.312 Å, β=104.6° and Z=2, refined to a conventional R factor of 0.034. The kinetics of the Hg2+- assisted chloride release from [CrCl(diamine)(3,3- tri)]ZnCl4 salts were measured spectrophotometrically (μ=1.0 M HClO4 or HNO3) over 15 K temperature ranges to give, in order, 104kHg (298.2 K) (M−1 s−1), Ea(kJ mol−1), ΔS# (J K−1 mol−1): en- (HClO4): 5.95, 78.1, -53; pn(HClO4); 5.24, 81.2; -44; tn(HClO4): 26.7, 85.6, -15; Me2tn(HClO4): 21.8, 78.6, -40; A-chxn(HNO3): 7.60, 81.0,-41; B-chxn(HNO3): 18.3, 56.8, -115. A ‘non-replaced ligand effect’ on the rate is observed for the first time in this series of homologous Cr(III) complexes. The kinetics of the thermal aquation (kH, 0.1 M HClO4) were measured titrimetrically for CrCl(diamine) (3,3-tri)2+ to give the following kinetic parameters: diamine=en: 107 kH (298.2)=5.34 s−1, Ea=99.2 kJ mol−1, ΔS#=-40 J K−1 mol-1; diamine =tn: 107 kH (298.2)=5.04 s−1, Ea= 82.8, ΔS#= -96.  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