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 205 毫秒
1.
The formation and dissociation of the binuclear complexes of FeIII withcis-[Co(en)2(RNH2)SalH]2+ [R=Me, Et and SalH=C6H4(OH)CO 2 ] were studied by a stopped-flow technique at 20–35°C, and I=1.0 mol dm−3 (ClO 4 ). The formation of the binuclear species, N5CoSalFe4+, involves reactions of the phenol form of the CoIII substrates with Fe(OH2) 6 3+ and Fe(OH2)5OH2+. The mechanism of reaction of Fe(OH2)5OH2+ is essentially Id, while that of Fe(OH2) 6 3+ appears to be Ia. The formation rate constant, k1, for Fe(OH2) 6 3+ /N5CoSalH2+ reaction decreases as the amine chain length increases, whereas the same (k2) for the Fe(OH2)5OH2+/N5CoSalH2+ reaction does not show any such trend. The binuclear species, N5CoSalFe4+, dissociates to yield a CoIII substrate and FeIII speciesvia a predominantly spontaneous dissociation path and a minor acid catalysed path which are relatively insensitive to the variation in size of the non-labile amine chain length.  相似文献   

2.
Rieske dioxygenases are metalloenzymes capable of achieving cis-dihydroxylation of aromatics under mild conditions using O2 and a source of electrons. The intermediate responsible for this reactivity is proposed to be a cis-FeV(O)(OH) moiety. Molecular models allow the generation of a FeIII(OOH) species with H2O2, to yield a FeV(O)(OH) species with tetradentate ligands, or {FeIV(O); OH.} pairs with pentadentate ones. We have designed a new pentadentate ligand, mtL42, bearing a labile triazole, to generate an “in-between” situation. Two iron complexes, [(mtL42)FeCl](PF6) and [(mtL42)Fe(OTf)2]), were obtained and their reactivity towards aromatic substrates was studied in the presence of H2O2. Spectroscopic and kinetic studies reflect that triazole is bound at the FeII state, but decoordinates in the FeIII(OOH). The resulting [(mtL42)FeIII(OOH)(MeCN)]2+ then lies on a bifurcated decay pathway (end-on homolytic vs. side-on heterolytic) depending on the addition of aromatic substrate: in the absence of substrate, it is proposed to follow a side-on pathway leading to a putative (N4)FeV(O)(OH), while in the presence of aromatics it switches to an end-on homolytic pathway yielding a {(N5)FeIV(O); OH.} reactive species, through recoordination of triazole. This switch significantly impacts the reaction regioselectivity.  相似文献   

3.
Properties indirectly determined, or alluded to, in previous publications on the titled isomers have been measured, and the results generally support the earlier conclusions. Thus, the common five‐coordinate intermediate generated in the OH?‐catalyzed hydrolysis of exo‐ and endo‐[Co(dien)(dapo)X]2+ (X=Cl, ONO2) has the same properties as that generated in the rapid spontaneous loss of OH? from exo‐ and endo‐[Co(dien)(dapo)OH]2+ (40±2% endo‐OH, 60±2% exo‐OH) and an unusually large capacity for capturing (R=[CoN3]/[CoOH][]=1.3; exo‐[CoN3]/endo‐[CoN3]=2.1±0.1). Solvent exchange for spontaneous loss of OH? from exo‐[Co(dien)(dapo)OH]2+ has been measured at 0.04 s?1 (k1, 0.50M NaClO4, 25°) from which similar loss from the endo‐OH isomer may be calculated as 0.24 s?1 (k2). The OH?‐catalyzed reactions of exo‐ and endo‐[Co(dien)(dapo)N3]2+ result in both hydrolysis of coordinated via an OH?‐limiting process =153 M ?1 s?1; =295 M ?1 s?1; KH=1.3±0.1 M ?1; 0.50M NaClO4, 25.0°) and direct epimerization between the two reactants =33 M ?1 s?1; =110 M ?1 s?1; 1.0M NaClO4, 25.0°). Comparisons are made with other rapidly reacting CoIII‐acido systems.  相似文献   

4.
The violet superoxo complex, [(H2O)4(OH)RhIII(O2)RhIII(OH)(H2O)4]3+, formed by treatment of (RhII)24+ with O2 in HClO4, is converted to a le? reduction product, the corresponding μ-peroxo complex, by the reductants I?, IrCl63?, and the trinuclear aquamolybdenum(III) cation, (MoIII)3. Each reaction is first-order in both redox partners, and the le? reduction by IrCl63? is followed by a much slower conversion to a peroxide-free complex. Among the rapid reductions of the superoxo derivative examined here and in a previous study, only that by IrCl63? is accelerated by increases in acidity; the rate law for this reaction features both an acid-independent and a [H+]-proportional component, the latter stemming from partial conversion of the oxidant to its conjugate acid (pKA < ?1.0). Rate laws for reductions by other metal-center reagents generally exhibit inverse-[H+] terms, reflecting deprotonation of the reductant. All reductions thus far observed involving this superoxo species appear to be outer-sphere. Treatment of acid-independent rate constants within the framework of the Marcus model, allows estimates of the self-exchange rate, k11, for the (RhIII)2-bound superoxo-peroxo couple. Because values of k11 calculated from the several reductions span a range of 104.5, reductions of the superoxo complex cannot be taken to conform satisfactorily to the Marcus treatment, being in this respect comparable to the systems VO(OH)+/2+, Mn2+/3+, Eu2+/3+, and Ti(OH)2+/3+, each of which exhibits similar divergences. The wide range of calculated self-exchange rates appears to invalidate an earlier suggestion that reduction of the superoxo complex by Fe2+ proceeds primarily through a bridged path. © 1994 John Wiley & Sons, Inc.  相似文献   

5.
A challenging objective of de novo metalloprotein design is to control of the outer coordination spheres of an active site to fine tune metal properties. The well-defined three stranded coiled coils, TRI and CoilSer peptides, are used to address this question. Substitution of Cys for Leu yields a thiophilic site within the core. Metals such as HgII, PbII, and AsIII result in trigonal planar or trigonal pyramidal geometries; however, spectroscopic studies have shown that CdII forms three-, four- or five-coordinate CdIIS3(OH2)x (in which x=0–2) when the outer coordination spheres are perturbed. Unfortunately, there has been little crystallographic examination of these proteins to explain the observations. Here, the high-resolution X-ray structures of apo- and mercurated proteins are compared to explain the modifications that lead to metal coordination number and geometry variation. It reveals that Ala substitution for Leu opens a cavity above the Cys site allowing for water excess, facilitating CdIIS3(OH2). Replacement of Cys by Pen restricts thiol rotation, causing a shift in the metal-binding plane, which displaces water, forming CdIIS3. Residue d -Leu, above the Cys site, reorients the side chain towards the Cys layer, diminishing the space for water accommodation yielding CdIIS3, whereas d -Leu below opens more space, allowing for equal CdIIS3(OH2) and CdIIS3(OH2)2. These studies provide insights into how to control desired metal geometries in metalloproteins by using coded and non-coded amino acids.  相似文献   

6.
In acetate buffer media (pH 4.5–5.4) thiosulfate ion (S2O32?) reduces the bridged superoxo complex, [(NH3)4CoIII(μ‐NH2,μ‐O2)CoIII(NH3)4]4+ ( 1 ) to its corresponding μ‐peroxo product, [(NH3)4CoIII(μ‐NH2,μ‐O2)CoIII(NH3)4]3+ ( 2 ) and along a parallel reaction path, simultaneously S2O32? reacts with 1 to produce the substituted μ‐thiosulfato‐μ‐superoxo complex, [(NH3)4CoIII(μ‐S2O3,μ‐O2)CoIII(NH3)4]3+ ( 3 ). The formation of μ‐thiosulfato‐μ‐superoxo complex ( 3 ) appears as a precipitate which on being subjected to FTIR shows absorption peaks that support the presence of Co(III)‐bound S‐coordinated S2O32? group. In reaction media, 3 readily dissolves to further react with S2O32? to produce μ‐thiosulfato‐μ‐peroxo product, [(NH3)4CoIII(μ‐S2O3,μ‐O2)CoIII(NH3)4]2+ ( 4 ). The observed rate (k0) increases with an increase in [TThio] ([TThio] is the analytical concentration of S2O32?) and temperature (T), but it decreases with an increase in [H+] and the ionic strength (I). Analysis of the log At versus time data (A is the absorbance of 1 at time t) reveals that overall the reaction follows a biphasic consecutive reaction path with rate constants k1 and k2 and the change of absorbance is equal to {a1 exp(–k1t) + a2 exp(–k2t)}, where k1 > k2.  相似文献   

7.
In the present study, poly(N-Isopropylacrylacrylamide-co-N-tertiarybutylacrylamide-co-hydroxyethylcrylamide) (NIPAM-co-NTBA-co-HEAAm) hydrogels are prepared with variation of molar ratio of hydrophilic HEAAm and hydrophobic NTBA. The prepared hydrogels are characterized with elemental analysis and Fourier transform infrared (FTIR) spectroscopy. The thermodynamics of swelling properties of poly(NIPAM-co-NTBA-co-HEAAm) hydrogels have also been discussed. The experimental C/N ratios are comparable with the theoretical value. The enthalpy change of mixing ∆Hmix, entropy change of mixing ∆Smix, free energy change of mixing ∆Gmix are determined for swelling of hydrogels at 25 °C. The value of total free energy of hydrogel swelling is found to be negative which confirms the lower critical solution temperature (LCST) exhibited in all hydrogels and the volume change transition shows the thermoresponsive behavior. The values of ∆Smix increase and ∆Gmix decrease with increasing amount of hydrophobic NTBA content in the hydrogels. The values of free energy change of elasticity (∆Gel) are found to be increased with increasing the hydrophobic NTBA content followed by decrease in swelling percentage. Also, the transition temperature of the hydrogel is found to be decreased with increasing the hydrophobic NTBA.  相似文献   

8.
The reaction of trans-[Cr(Salen)(OH2)2]+ with aqueous sulfite yields trans-[Cr(Salen)(OH2)(OSO2(SINGLEBOND)O)] (O-bonded isomer). The rate and activation parameter data for the formation of the sulfito complex are consistent with a mechanism involving rate-limiting addition of SO2 to the CrIII(SINGLEBOND)OH bond. The complex ions, trans-[(OH2)Cr(Salen)(OSO2(SINGLEBOND)O)], and trans-[(OH)Cr(Salen)(OSO2(SINGLEBOND)O)]2−, undergo reversible anation by NCS, N3, imidazole, and pyridine resulting in the formation of trans-[XCr(Salen)(OSO2(SINGLEBOND)O)](N+1)−(n=1 for X=N3,NCS, and 0 for X=imidazole and pyridine) predominantly via dissociative interchange mechanism. The labilizing action of the coordinated sulfite on the trans-CrIII-X bond in trans-[XCr(Salen)(OSO2)](n+1)− follows the sequence: NCSpyridine ca. N3 ca. imidazole. Data analysis indicated that the coordinated sulfite has little trans activating influence. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 373–384, 1998  相似文献   

9.
The influence of some amino acids—alanine, glycine, valine, glutamine, histidine, phenylalanine, and tryptophan—on the swelling behavior of acrylamide (AAm) hydrogel prepared by γ-radiation was investigated. Swelling experiments of AAm hydrogel were made in the universal buffer solutions and the amino acid solutions at certain pHs at 37°C. These selected pH values were pK1, pK2 and isoelectric point (pI) values such as ionization of α-carboxyl groups, ionization of α-amino groups, and the pIs of the amino acids, respectively. The swelling of AAm hydrogel increased when pH values of solutions were increased. The value of equilibrium swelling of AAm hydrogel in the solution of universal buffer was 880% at pH 10.0, whereas it was 670% at pH 2.0. The values of equilibrium swelling of AAm hydrogel in amino acid solutions were between 830 and 965% at pH 10.0, whereas they were between 635 and 775% at pH 2.0. The rate constant of swelling, initial swelling rate, theoretical maximum swelling, diffusional exponent, network parameter, and diffusion coefficient were calculated by swelling kinetics. Diffusions of the amino acid solutions into the hydrogel were generally found as non-Fickian in character. The diffusion coefficients of the hydrogel were between 0.91 × 10−6 and 2.41 × 10−6 cm2/s.  相似文献   

10.
The relationship among intrinsic surface reaction constant (K) in 1-pK model, point of zero net charge (PZNC) and structural charge density (σst) for amphoteric solid with structural charges was established in order to investigate the effect of σst on pK. The theoretical analysis based on 1-pK model indicates that the independent PZNC of electrolyte concentration (c) exists for amphoteric solid with structural charges. A common intersection point (CIP) should appear on the acid-base titration curves at different c, and the pH at the CIP is pHPZNC. The pK can be expressed as pK=-pHPZNC log[(1 2αPZNC)/(1-2αPZNC)], where αPZNC≡σst/eNANs, in which e is the elementary charge, NA the Avogadro‘s constant and Ns the total density of surface sites. For solids without structural charges, pK=-pHPZNC. The pK values of hydrotalcite-like compounds (HTlc) with general formula of [Mg1-xFex(OH)2](Cl,OH)x were evaluated. With increasing x, the pK increases, which can be explained based on the affinity of metal cations for H^- or OH^- and the electrostatic interaction between charging surface and H^- or OH^-.  相似文献   

11.
The mechanism of the reaction of trans‐ArPdBrL2 (Ar=p‐Z‐C6H4, Z=CN, H; L=PPh3) with Ar′B(OH)2 (Ar′=p‐Z′‐C6H4, Z′=H, CN, MeO), which is a key step in the Suzuki–Miyaura process, has been established in N,N‐dimethylformamide (DMF) with two bases, acetate (nBu4NOAc) or carbonate (Cs2CO3) and compared with that of hydroxide (nBu4NOH), reported in our previous work. As anionic bases are inevitably introduced with a countercation M+ (e.g., M+OH?), the role of cations in the transmetalation/reductive elimination has been first investigated. Cations M+ (Na+, Cs+, K+) are not innocent since they induce an unexpected decelerating effect in the transmetalation via their complexation to the OH ligand in the reactive ArPd(OH)L2, partly inhibiting its transmetalation with Ar′B(OH)2. A decreasing reactivity order is observed when M+ is associated with OH?: nBu4N+> K+> Cs+> Na+. Acetates lead to the formation of trans‐ArPd(OAc)L2, which does not undergo transmetalation with Ar′B(OH)2. This explains why acetates are not used as bases in Suzuki–Miyaura reactions that involve Ar′B(OH)2. Carbonates (Cs2CO3) give rise to slower reactions than those performed from nBu4NOH at the same concentration, even if the reactions are accelerated in the presence of water due to the generation of OH?. The mechanism of the reaction with carbonates is then similar to that established for nBu4NOH, involving ArPd(OH)L2 in the transmetalation with Ar′B(OH)2. Due to the low concentration of OH? generated from CO32? in water, both transmetalation and reductive elimination result slower than those performed from nBu4NOH at equal concentrations as Cs2CO3. Therefore, the overall reactivity is finely tuned by the concentration of the common base OH? and the ratio [OH?]/[Ar′B(OH)2]. Hence, the anionic base (pure OH? or OH? generated from CO32?) associated with its countercation (Na+, Cs+, K+) plays four antagonist kinetic roles: acceleration of the transmetalation by formation of the reactive ArPd(OH)L2, acceleration of the reductive elimination, deceleration of the transmetalation by formation of unreactive Ar′B(OH)3? and by complexation of ArPd(OH)L2 by M+.  相似文献   

12.
The redox reactions of thiosulfate with four iron(III) complexes having phenolate-amide-amine coordination, FeIII(L){L = 1,2-bis(2-hydroxybenzamido)ethane, L1; 1,3-bis(2-hydroxybenzamido)propane, L2; 1,5-bis(2-hydroxybenzamido)3-azapentane, L3; and 1,8-bis(2-hydroxybenzamido)3,6-diazaoctane, L4} have been investigated in 10% v/v MeOH + H2O and I = 0.3 mol dm−3. At constant pH (~ 4.8) and under pseudo-first order conditions of [S2O 3 2− ] the reaction obeyed the rate law : − d[FeIII(L)]/dt = k obs [FeIII(L)] + k obs where k obs denotes the observed rate constant of thiosulfate decomposition; k obs = a[S2O 3 2− ] + b[S2O 3 2− ] T 2 is valid for all the complexes, particularly at pH < 6, while k obs = [H+][S2O 3 2− ] T 2 is consistent with the rate law for thiosulfate decomposition proposed earlier. The rate data (k obs) were analysed on the basis of the reactivities of various species of FeIII(L) generated by the equilibrium protonation of the sec-NH of dien and trien spacer units resulting in the ring opening (for [FeIII(L3/L4)]), and acid base equilibrium of the aqua ligand bound to the iron(III) centre ([FeIII(L)(OH2) n ]). The redox activities, both for second and third order paths, show the ligand dependencies : L4<L3<L1<L2 conforming to the fact that the complexes tend to be less susceptible to electron transfer from S2O 3 2− with (i) the increase of the number of chelate rings, (ii) the decrease of overall charge, and (iii) the decrease of ring size offered by the amine moiety (from six membered to five membered one as for [FeIII(L1/L2)(OH2)2]+. There was no evidence for the formation of inner sphere thiosulfato complex, the possibility of the formation of the outer sphere ion-pairs, [Fe(L/HL)(OH2)n +/2+, S2O 3 2− ] with low equilibrium constant value may not be excluded. In view of this, the outer sphere electron transfer (ET) mechanism is the most likely possibility.  相似文献   

13.
Novel phosphorescent hydrogels have been explored by immobilizing an IrIII metal complex into the matrices of hydrogels. FTIR spectra demonstrate that the IrIII–PNaAMPS hydrogel is achieved by irreversible incorporation of positively charged [Ir(ppy)2(dmbpy)]Cl (ppy = 2‐phenylpyrine, dmbpy = 4,4′‐dimethyl‐2,2′‐bipyridine) into negatively charged poly(2‐acrylamido‐2‐methylpropane sulfonic acid sodium) (PNaAMPS) hydrogel via electrostatic interaction. The photoluminescent spectra indicate that the IrIII–PNaAMPS hydrogel exhibits stable phosphorescence. In vitro cultivation of human retinal pigment epithelial cells demonstrates the cytocompatibility of the IrIII–PNaAMPS hydrogel. This work herein represents a facile pathway for fabrication of phosphorescent hydrogels.  相似文献   

14.
The synthesis and crystal structure (at 100 K) of the title compound, Cs[Fe(C11H13N3O2S2)2]·CH3OH, is reported. The asymmetric unit consists of an octahedral [FeIII(L)2] fragment, where L2− is 3‐ethoxysalicylaldehyde 4‐methylthiosemicarbazonate(2−) {systematic name: [2‐(3‐ethoxy‐2‐oxidobenzylidene)hydrazin‐1‐ylidene](methylamino)methanethiolate}, a caesium cation and a methanol solvent molecule. Each L2− ligand binds through the thiolate S, the imine N and the phenolate O atoms as donors, resulting in an FeIIIS2N2O2 chromophore. The O,N,S‐coordinating ligands are orientated in two perpendicular planes, with the O and S atoms in cis positions and the N atoms in trans positions. The FeIII cation is in the low‐spin state at 100 K.  相似文献   

15.
The H2O‐exchange reaction on V(OH2), Mn(OH2), and Fe(OH2) has been reinvestigated with ab initio quantum‐chemical calculations that include electron correlation and hydration, whereby the second coordination sphere and the bulk solvent were treated as a dielectric continuum. In such models, activation entropies (ΔS) and also activation free energies (ΔG) are not available, since the second coordination sphere is not treated quantum chemically. Furthermore, no transition states for the dissociative interchange (Id) mechanism can be obtained, most likely also because of this approximation, and this limitation applies to the hexaaqua ions, but not, for example, to the pentaamines. Therefore, in cases where this model predicts that the dissociative (D) mechanism is the most favorable one, the question remains open as to whether the D or the Id mechanism operates. For the H2O adducts of the reactants, M(OH2)6⋅OH, two isomers were considered: in the first (A), the H2O molecule in the second coordination sphere forms a single H‐bond to one aqua ligand, and in the other (B), it is bound to two aqua ligands in a bridging mode. For each hexaaqua ion, associative (Ia or A) and dissociative mechanisms (D) were investigated. On the basis of reactant B, activation energies agreeing with experiment were obtained for the water exchange on the aqua ions of VII, MnII, and FeII. For the water exchange on V(OH2) and Fe(OH2), experimental and computational data suggest an a and a d activation, respectively, whereas for Mn(OH2), the activation energies, calculated for the a and d activations, are equal and, therefore, the mechanism can be attributed only via the comparison of the change of the sum of all MnII−O bond lengths during the activation process, ΔΣd(MnII−O), with the activation volume. The limitations in the attribution of substitution mechanisms on the basis of experimental activation volumes and activation energies computed with the present model are analyzed. The electronic structure of the heptacoordinated species, the transition state [V(OH2)5⋅⋅⋅(OH2)] and the intermediate Mn(OH2), are discussed.  相似文献   

16.
The ligand 2,6-bis(1-methylbenzimidazol-2-yl)pyridine (mbzimpy, 1 ) reacts with EuIII to give [Eu(mbzimpy)(NO3)3(CH3OH)] [ 4 ] whose crystal structure (EuC22H21N8O10, a = 7.658(3) Å, b = 19.136(2) Å, c = 8.882 Å, β = 104.07(1)°, monoclinic, P21, Z = 2) shows a mononuclear structure where EuIII is ten-coordinate by a meridional tridentate mbzimpy ligand, three bidentate nitrates, and one CH3OH molecule, leading to a low-symmetry coordination sphere around the metalion. Essentially the same coordination is found in the crystal structure of [Eu(obzimpy)(NO3)3] ( 8 ) (EuC35H45N8O9, a = 9.095(2) Å, b = 16.624(2) Å, c = 26.198(6) Å, β = 95.85(1)°, monoclinic, P21/c, Z = 4) obtained by reaction of 2,6-bis(1-octylbenzimidazol-2-yl)pyridine (obzimpy, 2 ) with EuIII. Detailed photophysical studies of crystalline [Ln(mbzimpy)(NO3)3(CH3OH)] and [Ln(obzimpy)(NO3)3] complexes (Ln = Eu, Gd, Tb, Lu) show that 1 and 2 display 1ππ* and 3ππ* excited states very similar to those observed in 2,2′:6′,2″-terpyridine, leading to efficient ligand to LnIII intramolecular energy transfer. Spectroscopic results show that an extremely efficient mbzimpy-to-EuIII transfer occurs in [Ln(mbzimpy)(NO3)3(CH3OH)] and in the case of TbIII, a TbIII-to-mbzimpy back transfer is also implied in the deactivation process. The origin of these peculiar effects and the influence of ligand design by going from mbzimpy to obzimpy are discussed. 1H-NMR and luminescence data indicate that the structure found in the crystal is essentially maintained in solution.  相似文献   

17.
Dispersion of hydrogels of FeIII, InIII, AlIII, CrIII, TiIV, and ZrIV oxohydroxides in an aqueous medium was studied by sedimentation analysis. The hydrogel dispersion depends on the metal nature, pH of precipitation, and suspension concentration. The systems are predominantly polydisperse, and the gel particle size ranges from 2 to 140 μm. The data obtained suggest that the gel particles are formed by three-dimensional networks consisting of polymeric chains metal-oxygen and contain cavities filled with water. __________ Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 5, pp. 1083–1088, May, 2005.  相似文献   

18.
S.-L. Wu  F.-C. Lu 《Liquid crystals》2013,40(11):1517-1523
Three homologous series of chiral swallow-tailed compounds, alkyl (S)-2-{6-[4-(4′-alkoxyphenyl)benzoyloxy]-2-naphthyl}propionates, (S)HNP(p,n,q) derived from (S)-2-(6-hydroxy-2-naphthyl)propionic acid in conjugation with a variety of swallow-tailed groups, attached to the external side of the chiral centre, have been synthesized and their mesomorphic and electro-optical properties studied. Both (S)HNP(p,1,2) and (S)HNP(p,1,3) exhibited an enantiotropic antiferroelectric SmC*A phase. This implys that the swallow-tailed groups in the molecules favour zigzag pairing of the molecules in the smectic phase. The maximum P S values of compounds (S)HNP(p,1,2) in the antiferroelectric phase were measured in the range 21–30 nC cm-2; those of compounds (S)HNP(p,1,3) were in the range 15–23 nC cm-2, indicating that these chiral compounds possess low polarity. The electro-optical response of the compounds in the antiferroelectric SmC*A phase displayed thresholdless V-shaped switching.  相似文献   

19.
The cobalt substituted polyoxotungstate [Co6(H2O)2(α-B-PW9O34)2(PW6O26)]17− ( Co6 ) displays fast electron transfer (ET) kinetics to photogenerated RuIII(bpy)33+, 4 to 5 orders of magnitude faster than the corresponding ET observed for cobalt oxide nanoparticles. Mechanistic evidence has been acquired indicating that: (i) the one-electron oxidation of Co6 involves Co(II) aquo or Co(II) hydroxo groups (abbreviated as Co6(II) −OH 2 and Co6(II) −OH, respectively, whose speciation in aqueous solution is associated to a pKa of 7.6), and generates a Co(III)−OH moiety ( Co6(III) −OH), as proven by transient absorption spectroscopy; (ii) at pH>pKa, the Co6(II) −OH→RuIII(bpy)33+ ET occurs via bimolecular kinetics, with a rate constant k close to the diffusion limit and dependent on the ionic strength of the medium, consistent with reaction between charged species; (iii) at pH <pKa, the process involves Co6(II) − OH2 → Co6(III)−OH transformation and proceeds via a multiple-site, concerted proton electron transfer (CPET) where water assists the transfer of the proton, as proven by the absence of effect of buffer base concentrations on the rate of the ET and by a H/D kinetic isotope in a range of 1.2–1.4. The reactivity of water is ascribed to its organization on the surface of the polyanionic scaffold through hydrogen bond networking involving the Co(II)−OH2 group.  相似文献   

20.
DNA‐tethered poly‐N‐isopropylacrylamide copolymer chains, pNIPAM, that include nucleic acid tethers have been synthesized. They are capable of inducing pH‐stimulated crosslinking of the chains by i‐motif structures or to be bridged by Ag+ ions to form duplexes. The solutions of pNIPAM chains undergo crosslinking at pH 5.2 or in the presence of Ag+ ions to form hydrogels. The hydrogels reveal switchable hydrogel‐to‐solution transitions by the reversible crosslinking of the chains at pH 5.2 and the separation of the crosslinking units at pH 7.5, or by the Ag+ ion‐stimulated crosslinking of the chains and the reverse dissolution of the hydrogel by the cysteamine‐induced elimination of the Ag+ ions. The DNA‐crosslinked hydrogels are thermosensitive and undergo reversible temperature‐controlled hydrogel‐to‐solid transitions. The solid pNIPAM matrices are protected against the OH? or cysteamine‐stimulated dissociation to the respective polymer solutions.  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