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 140 毫秒
1.
A series of α-hydroxyalkylferrocenes (ferrocenemethanol [FcCH2OH], α-hydroxyethylferrocene, and 1,1′-ferrocenedimethanol) have been shown to incorporate into highly oxidized films of poly(aniline) [PAn]. Surface immobilization is readily evident in cyclic voltammetric studies by the rapid growth of a pair of sharp redox waves at E1/2=0.455 V versus SCE in the case of FcCH2OH. A general mechanism for the attack of α-hydroxyalkylferrocenes on oxidized PAn is inferred from the electrochemical data.  相似文献   

2.
A pyrocatechol sulfonephthalein- (PS-) modified glassy carbon (PS/GC) electrode has been prepared by adsorption of PS on a glassy carbon electrode surface. Cyclic voltammograms of the PS/GC electrode indicate the presence of a couple of well-defined redox peaks, and the formal potential shifts in the negative direction with increasing solution pH. The relation between formal potential,E0′, and solution pH can be fit to the equationE0′(mV) = −51.4 pH + 538.7. The PS/GC electrode shows high electrocatalytic activity toward ascorbic acid oxidation, with an overpotential ca. 380 mV less than that of the bare electrode and a drastic enhancement of the anodic currents. The electrocatalytic reaction rate constant (k), which was decreased with increasing concentration of H2A, was determined using rotating disk electrode measurements. The values ofkwas also affected by the solution pH. The electrode can also separate the electrochemical responses of ascorbic acid and dopamine. The separation between the anodic peak potentials of ascorbic acid and dopamine is more than 50 mV by the differential pulse voltammetry.  相似文献   

3.
《Polyhedron》1995,14(23-24)
New complexes of bivalent nickel with isopropylxanthates and nitrogen-donor ligands of composition [Ni(Prixa)2(L)], [Ni(Prixa)2(L1)2], [Ni(L2)2](Prixa)2, and [Ni(L3)3] (Prixa)2 have been synthesized, where Prixa = i-C3H7OCS2, L = 1,2-diaminopropane (1,2-pn), N,N,N′,N′=tetramethylethylenediamine (tmen) or 4,4′-bipyridine (4,4′-bipy), L1 = pyridine (py), L2 = diethylenetriamine (dien) and L3 = ethylenediamine (en), 1,2-diaminopropane or 1,10-phenanthroline (phen). The compounds have been characterized by elemental analysis, IR and UV-vis spectroscopy, magnetochemical measurements, molar conductivity and thermal analysis. The compounds containing the complex cation have been one-electron irreversibly oxidized using cyclic voltammetry. The crystal and molecular structures of [Ni(Prixa)2(tmen)] and [Ni(phen)3](Prixa)2 have been elucidated.  相似文献   

4.
The reaction of the heteroleptic Nd(III) iodide, [Nd(L′)(N″)(μ-I)] with the potassium salts of primary aryl amides [KN(H)Ar′] or [KN(H)Ar*] affords heteroleptic, structurally characterised, low-coordinate neodymium amides [Nd(L′)(N″)(N(H)Ar′)] and [Nd(L′)(N″)(N(H)Ar*)] cleanly (L′ = t-BuNCH2CH2[C{NC(SiMe3)CHNt-Bu}], N″ = N(SiMe3)2, Ar′ = 2,6-Dipp2C6H3, Dipp = 2,6-Pri2C6H3, Ar* = 2,6-(2,4,6-Pri3C6H2)2C6H3). The potassium terphenyl primary amide [KN(H)Ar*] is readily prepared and isolated, and structurally characterised. Treatment of these primary amide-containing compounds with alkali metal alkyl salts results in ligand exchange to give alkali metal primary amides and intractable heteroleptic Nd(III) alkyl compounds of the form [Nd(L′)(N″)(R)] (R = CH2SiMe3, Me). Attempted deprotonation of the Nd-bound primary amide in [Nd(L′)(N″)(N(H)Ar*)] with the less nucleophilic phosphazene superbase ButNP{NP(NMe2)3}3 resulted in indiscriminate deprotonations of peripheral ligand CH groups.  相似文献   

5.
A transition metal complex as an electrochemical probe of a DNA sensor must have an applicable redox potential, high binding affinity and chemical stability. Some complexes with the dipyrido[3,2-a:2′,3′-c]phenazine (DPPZ) ligand have been reported to have high binding affinity for DNA. However, it was difficult to detect the targeted DNA electrochemically using these complexes because of the relatively high redox potential. In this work, a combination of bipyridine ligands with functional groups (---NH2, ---CH3 and ---COOH) and the DPPZ ligand were studied. The introduction of electron-donating groups was effective for controlling the redox potential of the DPPZ-type osmium complex. The [Os(DA-bpy)2DPPZ]2+ complex (DA-bpy; 4,4′-diamino-2,2′-bipyridine) had a lower half-wave potential (E1/2) of 147 mV (vs. Ag AgCl) and higher binding affinity with DNA {binding constant, K=3.1×107 M−1 in 10 mmol dm−3 Tris–HCl buffer with 50 mmol dm−3 NaCl (pH 7.76)} than those of other complexes. With the single stranded DNA (ssDNA) modified gold electrode, the hybridization signal (ΔI) of the [Os(DA-bpy)2DPPZ]2+ complex was linear in the concentration range of 1.0 pg ml−1–0.12 μg ml−1 for the targeted DNA with a regression coefficient of 0.999. The detection limit was 0.1 pg ml−1.  相似文献   

6.
The electrooxidation of vitamin D2 (VD2) was studied by cyclic voltammetry and in situ circular dichroic (CD) spectroelectrochemistry for the first time. The mechanism of electrooxidation and some useful kinetic and adsorption parameters were obtained. The results showed that the oxidation of VD2 in ethanol solution is an irreversible diffusion controlled process following a weak adsorption of the electroinactive product at a glassy carbon electrode, which blocks the electrochemical reaction. The electrooxidation occurs mainly at the triene moieties of the VD2 molecule. The CD spectroelectrochemical data were treated by the double logarithm method together with nonlinear regression, from which the formal potential E0=1.08 V, αn=0.245, the standard electrochemical rate constant k0=4.30(±0.58)×10−4 cm s−1, and the adsorption constant β=1.77(±0.25) were obtained.  相似文献   

7.
Speeds of sound, uijk, of 1,3-dioxolane or 1,4-dioxane (i) + water (j) + formamide or dimethylformamide (k) ternary mixtures and of their binary subsystems, uij, of 1,3-dioxolane or 1,4-dioxane (i) + formamide or dimethylformamide (j), and water (i) + formamide or dimethylformamide (j) have been measured over the entire composition range at 308.15 K. The experimental data have been used to evaluate the excess isentropic compressibilities of binary (κsE)ij and ternary (κsE)ijk mixtures using their densities calculated from molar excess volume data. The Moelwyn-Huggins concept [M. L. Huggins, Polymer 12, 389 (1971)] of interaction between the surfaces of components of a binary mixture has been employed to evaluate the excess isentropic compressibilities (using the concept of connectivity parameter of third degree of a molecule, 3ξ, which in turn depends on its topology) of binary mixtures, and this method has been extended to predict excess compressibilities of ternary mixtures. Values of (κsE)ij and (κsE)ijk have also been calculated by the Flory theory. It was observed that (κsE)ij and (κsE)ijk predicted by the Moelwyn-Huggins approach compare well with calculated and experimental values.  相似文献   

8.
Speed of sound data, uijk, of 1,3-dioxolane or 1,4-dioxane(i) + water(j) + propan-1-ol or propan-2-ol(k) ternary mixtures and their sub-binary mixtures, uij, of 1,3-dioxolane or 1,4-dioxane(i) + water or propan-1-ol or propan-2-ol(j) and water(i) + propan-1-ol or propan-2-ol(j) mixtures have been measured over the entire composition range at 308.15 K. Isentropic compressibility changes of mixing, (κsE)ij and (κsE) ijk, for the binary and ternary mixtures have been determined by employing the observed speeds of sound data and densities (calculated from their molar excess volumes data). The (κsE) ij and (κsE) ijk values have also been predicated by the graph theoretical approach and the Flory theory. It has been observed that (κsE) ij and (κsE) ijk predicted by the graph theoretical approach compare well with their corresponding experimental values.  相似文献   

9.
Ag+-assisted dechlorination of blue cis-trans-cis Ru(R-aai-R′)2Cl2 followed by the reaction with chloranilic acid (H2CA) in the presence of Et3N, gives a neutral mononuclear violet complex [Ru(R-aai-R′)2(CA)]. [R-aai-R′=p-R-C6H4—N=N—C3H2—NN, abbreviated as an N,N′ chelator where N(imidazole) and N(azo) represent N and N′, respectively; R = H (a), OMe (b), NO2 (c) and R′= Me (4), Et(5), Bz(6)]. All the complexes exhibit strong intense MLCT transitions in the visible region and weak broad bands at higher wavelength (>700 nm). Visible transitions (580–595 nm) show a negative solvatochromic effect. The cyclic voltammograms show two quasireversible to irreversible couples positive to SCE and are due to CA/CA2− (1.2–1.35 V) and Ru(III)/Ru(II) (1.6–1.8 V) redox processes. Three couples, negative to SCE, are assigned to CA2−/CA3− (−0.2 to −0.3 V), and azo reductions (−0.5 to −0.7, −0.8 to −0.9 V) of the chelated R-aai-R′.  相似文献   

10.
The title compound, cobalt 4′,7-diethoxylisoflavone-3′-sulfonate([Co(H2O)6](X)2⋅8H2O, X = C19H17O4SO3) was synthesized and its structure was determined by single-crystal X-ray diffraction analysis. It crystallizes in the triclinic space group P-1 with cell parameters a = 9.026(3) Å, b = 16.431(5) Å, c = 18.195(6) Å, α = 72.289(4), β = 87.498(4), γ = 82.775(5), V = 2550.1(13) Å−3, Dc = 1.419 Mg m−3, and Z = 2. The results show that the title compound consists of one cobalt cation, six coordinated water molecules, eight lattice water molecules, and two 4′,7-diethoxylisoflavone-3′-sulfonate anions, C19H17O4SO3. Two anions have different conformations. Twelve H atoms of six coordinated water molecules, as donors, form hydrogen bonds with four oxygen atoms of sulfo-groups of two anions and eight oxygen atoms of eight lattice water molecules. In addition, π < eqid1 > ⋅ < eqid2 > π stacking interactions exist in the crystal structure, which together with hydrogen bonds lead to supramolecular formation with a three-dimensional network.  相似文献   

11.
N-bromosuccinimide (NBS) oxidation of some substituted benzhydrols (4-Cl-, 4-Me-, 4-OMe-, 4-NO2−, 4,4′-dimethyl-, 4,4′-dichloro-, 4,4′-dimethoxy-, and 4-methyl-2-nitro-) yields corresponding benzophenones in the presence of Hg(OAc)2. The reaction is first-order in [NBS], [substrate], and [H+]. A study on the primary kinetic hydrogen isotope effect and solvent isotope effect suggest that C-H and O-H stretching frequencies are affected in the transition state. Activation parameters for the rate-determining step have been evaluated. The results are in accord with the linear free energy relationship (LFER). The linear plot of log(k obs/k 0) vs. σ+ (ρ = −0.69)and Bunnett plots support the existence of proton-transfer in the rate-determining step. Rate behavior in different solvent compositions suggests dipolar-dipolar interaction in the absence of acid and ion-dipolar interaction in the presence of acid. Two different mechanisms have been suggested: cyclic transition state with unprotonated N-bromosuccinimide (NBS) in the absence of acid and noncyclic transition state with protonated NBS in the presence of acid.__________From Kinetika i Kataliz, Vol. 46, No. 3, 2005, pp. 360–365.Original English Text Copyright © 2005 by Hiran, Malkani, Rathore.This article was submitted by the authors in English.  相似文献   

12.
The complex [Ru(II)(dcbpyH2)(bdmpp)NCS](PF6) (1) (where dcbpyH2 is 2,2′-bipyridine-4,4′-dicarboxylic acid, bdmpp is 2,6-bis(3,5-dimethyl-N-pyrazoyl)pyridine,) is synthesized and characterized extensively by 1H NMR and 13C NMR 1D and 2D, mass spectroscopy, cyclic voltammetry, electronic absorption spectroscopy and IR. The half-wave potential of the Ru(II)/Ru(III) redox couple was measured at E1/2=+0.795 V versus Ag/AgCl in CH3CN. The complex presents three intense metal-to-ligand charge transfer (MLCT) (dM→πL*) absorption bands centered at 383 (=21 300 M−1 cm−1), 432 (=22 400 M−1 cm−1) and 475 nm (=23 400 M−1 cm−1), respectively. The absorbance is extremely strong between 400 and 500 nm and even at 620 nm, the extinction coefficient is still high (=3768 M−1 cm−1). The strong π-acceptor property of the trans-isothiocyanate ligand compared with the Cl ligand is probably the cause of the blue-shift observed in complex 1. These properties make the complex potentially promising for the photosensitization process. The incorporation of TiO2 photoelectrodes derivatized with this complex into a solar cell using a composite polymer/inorganic oxide solid-state electrolyte confirmed its sensitizing ability. Incident monochromatic photon-to-current conversion efficiency (IPCE) values of about 30% and overall energy conversion efficiency (η) of 1.7% were obtained.  相似文献   

13.
The new Cs3Mo6Br13O oxybromide, synthesized by solid-state chemistry, crystallizes in the trigonal system (Rc space group; a = 15.5784(2) Å, c = 19.5103(5) Å, V = 4100.5(1) Å3 and Z = 6). It is based on a [Mo6L14] unit that contains an unprecedented μ3 face-capping oxygen. The crystal structure determined by single crystal X-ray diffraction is built up from discrete face-capped [Mo6Bri6Li2Bra6]3– (L = 0.5 O + 0.5 Br) anionic units in which two inner positions are randomly occupied by one bromine and one oxygen whereas the other ligand positions are fully occupied by bromine. The cesium cations randomly occupy two close crystallographic positions generated by the A-B-C-A-B-C′ close-packed stacking of the units. The cesium site occupancy is related to the random distribution of oxygen and bromine on the Li inner positions. To cite this article: K. Kirakci et al., C. R. Chimie 8 (2005).  相似文献   

14.
The photochemical reactivity, photophysical properties and redox behavior of the complexes trans,cis-[Ru(X)(X′)(CO)2(α-diimine)] and their derivatives are strongly dependent on the complex geometry, the nature and electronic properties of the α-diimine ligand and, most importantly, on the axial ligands X and X′ (alkyl, halide, phosphine, donor solvent, etc.). This paper deals mainly with comparison of reduction pathways for several different types of the trans,cis-[Ru(X)(X′)(CO)2(α-diimine)] complexes, also presenting some new results in this field. An equally important goal has been the comparison and discussion of the photo- and redox reactivity of these complexes from the viewpoint of the frontier orbitals involved and character of the Ru---X/X′ bonding.  相似文献   

15.
The role of fluorinated β-diketones, their tautomers (keto–enols) and their derivatives as reagents towards λ3P compounds is reviewed, including 2-trifluoroacetyl phenols, possessing formally a keto–enol system, and their derivatives. In an ‘insertion’ reaction phosphine and the keto–enol tautomers of 1,1,1,5,5,5-hexafluoro- and 1,1,1-trifluoropentan-2,4-dione furnished primary (S) or (R) α-hydroxy phosphines, whose enol functions probably isomerized the corresponding keto compounds. Further addition and isomerisation furnished 1,3α,5,7β-tetrakis(trifluoromethyl)-2-phospha-6-oxa-9-oxabicyclo[3.3.1]-nonan-3β,7α-diol and 1,7-trifluoromethyl-3,5-methyl-2,4,8-trioxa-6-phophaadamantane, exclusively one diastereomer in each case. The main mechanistic feature of these reactions is a consecutive diastereoselective hemiketal cyclization. 1,1,1,5,5,5-Hexafluoro- and 1,1,1-trifluoropentan-2,4-dione, as well as 2-trifluoroacetyl phenol and its imino derivatives reacted diastereospecifically with phosphonous acid dichlorides, RPCl2 to give in a concerted mechanism thermally stable tricyclic λ5σ5P phosphoranes containing two five-membered rings and one six-membered ring. Surprisingly, the two CF3 groups bonded to an sp3-hybridized carbon were in a cisoid arrangement having closest non-bonding FF distances of 301.4 or 273.5 pm. These findings reflect the ‘through space’ F---F coupling constants of the tricyclic phosphoranes (JFF=4.0–7.0 Hz), in solution. 4,4,4-Trifluoro-3-hydroxy-1-phenyl-butan-1-one and methyl or phenyl phosphonous acid dichlorides gave similar tricyclic phosphoranes decomposing at ambient temperature to furnish 1,2λ5σ4-oxaphospholanes and (E)-1,1,1-trifluoro-4-phenyl-but-2-en-4-one. Dialkylphosphites and 1,1,1,5,5,5-hexafluoropentan-2,4-dione reacted to give either the (Z)-enol phosphonates or the respective γ-ketophosphonates from which in two cases four diastereomeric 2-oxo-2,5-dialkoxy-3,5-bis(trifluoromethyl)-3-hydroxy-1,2λ5σ4-oxa-phospholanes were obtained. 2-Trifluoroacetyl cyclohexanone, 4,4,4-trifluoro-3-trimethylsiloxy-1-phenylbutan-1-one, 1-benzoyl-2-trifluormethyloxirane, 1-benzoyl-2-trifluoro-methylaziridine, 2-trifluoroacetyl-1-trimethylsiloxybenzene and (trifluoroacetyl-1-phenyl) diethyl phosphate reacted with tris(trimethylsilyl) phosphite to give functionalized α-trimethylsiloxy phosphonates, which could easily be transferred into the respective phosphonic acids. In the case of an oxirane and an aziridine ketone no ring cleavage was observed. For 1,1′-(2-hydroxy-5-methyl-m-phenylene)-bis-ethanone and 1,1′-(2-trimethylsiloxy-5-methyl-m-phenylene)-bis-ethanone benzoxaphospholanes were obtained. Trialkyl phosphites and 1,1,1,5,5,5-hexafluoropentan-2,4-dione furnished cyclic phosphoranes containing the 3-hydroxy-3,5-bis(trifluoromethyl)-1,2λ5σ5-oxaphospholene structural element, stable at ambient temperature only in the case of one cyclic phosphite precursor. (E)-1,1,1-Trifluoro-4-phenyl-but-2-en-4-one and trimethylphosphite reacted to form 1,2λ5σ5-oxaphosphol-4-ene as the sole product. Results similar to the reaction of 1,1′-(2-hydroxy-5-methyl-m-phenylene)-bis-ethanone with diethyltrimethylsilylphosphite were obtained for trimethylphosphite and 2-trifluoroacetyl phenol where a deoxygenated phosphorane was found, easily hydrolyzed to give the respective phosphonic acid. With dialkylisocyanato phosphites and the keto components, 1,1,1,5,5,5-hexafluoro- and 1,1,1-trifluoropentan-2,4-dione, 4,4,4-trifluoro-1-phenyl-1,3-butandione, 2-trifluoroacetyl cyclohexanone, 2-trifluoroacetyl phenol and 1,1′-(2-hydroxy-5-methyl-m-phenylene)-bis-ethanone reacted in a ‘double’ cycloaddition to form bicyclic phosphoranes containing the 4,8-dioxa-2-aza-1λ5σ5-phosphabicyclo[3.3.0]-oct-6-en-3-one ring system; for the imino derivatives of 2-trifluoroacetyl phenol a corresponding 8-oxa-2,4-diaza- system was generated. For (E)-1,1,1,5,5,5-hexafluoro-4-trimethylsiloxy-3-penten-2-one however, a cyclic spiroimino phosphorane was obtained which underwent a [2+2] cyclodimerization to form a diazadiphosphetidine. Dimethylpropynyl phosphonite and 1,1,1,5,5,5-hexafluoropentan-2,4-dione yielded diastereoselectively a bisphosphorane, namely 1,4-bis(trifluoromethyl)-3,6-dioxa-2,2,7,7-tetramethoxy-2,7-di(1-propynyl)-2,7-diphosphabicyclo[2.2.1] heptane. When trimethylsilanyl–phosphenimidous acid bis-trimethylsilanyl–amide, Me3SiN=PN(SiMe3)2, was allowed to react with 1,1,1,5,5,5-hexafluoro- and 1,1,1-trifluoropentan-2,4-dione, (E)-1,1,1,5,5,5-hexafluoro-4-trimethylsiloxy-3-penten-2-one, 2-trifluoroacetyl cyclopentanone, 2-trifluoroacetyl phenol and its imino derivatives, 2-imino-1,2λ5σ4-oxaphospholenes were found containing two diastereomers in each case, which added hexafluoroacetone across the P=N bond to give 1,3,2λ5σ5-oxazaphosphetanes.  相似文献   

16.
With bis(alkynyl) titanocene ligands acting as chelating moieties for low-valent transition-metal heterobimetallic Ti(IV)---M(I) (M=Cu, Ag), complexes of the type {[Ti](CCSiMe3)}MX ([Ti]=(η5-C5H4SiMe3)2Ti; X=Cl; Br; I) were prepared. On reacting such complexes with {[Ti](CCSiMe3)2}MX′ (X′=OClO3) tetrametallic species of the type [{[Ti](CCSiMe3)2}M---X---M{(Me3SiCC)2[Ti]}]+ are formed. Their redox electrochemistry was studied with cyclic voltammetry (CV); for comparison, data of related bimetallic complexes were obtained. Results indicate a strong intramolecular interaction between the Group 11 metal held in place by the organometallic {[Ti](CCSiMe3)2} π-tweezers.  相似文献   

17.
The rare-earth (R) platinum aluminidesRPtAl crystallize in the orthorhombic TiNiSi-type structure (space group Pnma,Z=4), where magnetic rare-earth atoms form a network of chains parallel to thea-axis and parallel to theb-axis. Magnetic structures and phase transitions ofRPtAl (R=Ce, Pr, Nd) compounds were investigated by systematic measurements of magnetic susceptibility, specific heat, and neutron diffraction on polycrystalline samples. The results reveal a large magnetocrystalline anisotropy and magnetic structures that are dominated by a ferromagnetic component parallel to one of the two chain directions: thea-axis for CePtAl and PrPtAl and theb-axis for NdPtAl. The complex magnetism of CePtAl with three successive magnetic phase transitions (TC=5.9 K,T2=4.3 K,T3=2.5 K) and two coexisting propagation vectors (k1=0 forTTC, k2i=[0, 0.46, 0] forT2TTC, k2=[0, 1/2, 0] forTT2) is confirmed to be exceptional amongRPtAl compounds. PrPtAl has a nonmagnetic crystalline-electric field (CEF) ground-state singlet separated by 21 K from the first-excited state CEF singlet and magnetic exchange interactions are strong enough to induce long-range magnetic order (Curie temperatureTC=5.8 K, propagation vector k1=0, magnetic group Pnm′a′, ordered saturation momentm1=1.00(7)μB). NdPtAl is a simple ferromagnet (TC=19.2 K, k1=0, Pn′ma′,m1=2.08(4)μB).  相似文献   

18.
Two polymorphs of zero-dimensional zinc phosphate with the formula, 0[Zn(2,2′-bipy)(H2PO4)2], have been synthesized employing hydrothermal technique and their structure determined by single crystal X-ray diffraction. Both the structures consists of ZnO3N2 distorted trigonal-bipyramidal and PO2(OH)2 tetrahedral units linked through their vertices giving rise to a zero-dimensional molecular zinc phosphate. The structures are stabilized by extensive hydrogen bond interactions between zero-dimensional monomers. The structures display subtle differences in their packing created by hydrogen bond interactions. Crystal data: polymorph I, triclinic, space group (No. 2), , , , α=67.32(3)°, β=81.67(3)°, γ=69.29(3)°, , Z=2; polymorph II, triclinic, space group (No. 2), , , , α=97.37(2)°, β=100.54(2)°, γ=100.98(2)°, , Z=2.  相似文献   

19.
The l,2-bis(sulphapyridyl)oxamide ligand [L] and its complexes with FeIII, CoII, CuII and ZnII chloride were synthesized and characterized by elemental analyses, i.r., n.m.r., e.p.r. and u.v.–vis. spectroscopy and molar conductance measurements. Spectroscopic studies show that all the complexes are octahedral and covalent. The electrochemical behaviour of the CoII complex was monitored by cyclic voltammetry in a buffer/DMF solution (95:5). The E 0 values –0.622 and –0.502 V reveal a reversible one electron redox wave attributed to a CoII/CoI redox couple at a scan rate of 0.1 V s–1. The interaction of the CoII complex with bovine milk casein (BMC) was studied at the same scan rate, which reveals a strong binding as the E 0 values shift to more negative potential (E 0 = –0.908 and –0.703 V). The cyclic voltammograms of the CoII complex bound by BMC were recorded at different pH's. The plot of E 0 versus pH showed that E 0 values are maximal at pH 7.4 indicating good interaction between the BMC and the CoII complex which is further confirmed by kinetic data. The kinetic studies of the CoII complex bound to BMC was monitored in phosphate buffer solution at different pH's by spectrophotometry. The absorbance changes were monitored at 278 nm ( max for BMC) with respect to time and pseudo-first-order rate constants, K obs, were obtained from the slope and intercept of the straight line using the least squares regression method. The plot of absorbance versus time at different pH's was linear up to 80% completion of the reaction. The pH-rate profile data reveals that the reactions are pH dependent.  相似文献   

20.
Electrochemistry of a mixture of hydrofullerenes C70H36—46 composed of C70H36, C70H38, C70H44, and C70H46 (50, 20, 14, and 15%, respectively) was studied by cyclic voltammetry in THF and CH2Cl2 in the –43—–13 °C temperature range. Two cathodic peaks, namely, one-electron reversible (E° = –3.16 V (Fc0/+), Fc is ferrocene) and irreversible (E p = –3.37 V (Fc0/+)) were observed for this mixture in THF. The irreversible broad oxidation peak (E p = 1.22 V (Fc0/+)) was observed in CH2Cl2. The reversible reduction peak (E° = –3.16 V) and irreversible oxidation peak (E p = 1.22 V) were attributed to the most stable hydrofullerene C70H36. The irreversible reduction (E p = –3.37 V) and oxidation (E p = 1.22 V) peaks were attributed to hydrofullerenes C70H44—46 with a higher degree of hydrogenation. The values of an electrochemical gap, which is an analog of the energy gap (HOMO—LUMO), are 4.38 and 4.59 V for C70H36 and C70H44—46, respectively, and indicate that these hydrofullerenes are sufficiently hard molecules with low reactivity in redox reactions.  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