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 749 毫秒
1.
The mechanism of the oxygen evolution reaction (OER) by catalysts prepared by electrodepositions from Co(2+) solutions in phosphate electrolytes (Co-Pi) was studied at neutral pH by electrokinetic and (18)O isotope experiments. Low-potential electrodepositions enabled the controlled preparation of ultrathin Co-Pi catalyst films (<100 nm) that could be studied kinetically in the absence of mass transport and charge transport limitations to the OER. The Co-Pi catalysts exhibit a Tafel slope approximately equal to 2.3 × RT/F for the production of oxygen from water in neutral solutions. The electrochemical rate law exhibits an inverse first order dependence on proton activity and a zeroth order dependence on phosphate for [Pi] ≥ 0.03 M. In the absence of phosphate buffer, the Tafel slope is increased ~3-fold and the overall activity is greatly diminished. Together, these electrokinetic studies suggest a mechanism involving a rapid, one electron, one proton equilibrium between Co(III)-OH and Co(IV)-O in which a phosphate species is the proton acceptor, followed by a chemical turnover-limiting process involving oxygen-oxygen bond coupling.  相似文献   

2.
Three new isomorphic coordination polymers of Co(2+), Zn(2+) ions with flexible multicarboxylic acid ligand of the cis,cis,cis-1,2,3,4-cyclopentanetetracarboxylic acid (H(4)L), [Co(4)L(2)(H(2)O)(8)]·3H(2)O (1), [Zn(4)L(2)(H(2)O)(8)]·3H(2)O (2) and [Co(0.8)Zn(3.2)L(2)(H(2)O)(8)]·3H(2)O (3), have been synthesized under hydrothermal conditions and by means of controlling the pH of the reaction mixtures (with an initial pH of 6.0 for 1, 4.0 for 2, and 5.0 for 3, respectively). In the crystal of 1, two crystallographically different Co(2+) ions (Co1 and Co2) form a negatively-charged coordination polymeric chain, which contains a centrosymmetric, linear, trinuclear Co(2+) cluster (Co(3)L(2)) subunit; another crystallographically independent Co(2+) ion (Co3) coordinated to six water molecules acts as a counter ions to link the neighboring coordination polymeric chains via intermolecular H-bond interactions. The Co(2+) ions in 1 were completely and partially replaced by Zn(2+) ions to give 2 and 3, respectively. Complex 3 shows a novel molecular alloy nature, due to the random distributions of the Co(2+) and Zn(2+) ions. Three isomorphic complexes exhibit distinct thermal decomposition mechanisms. The deprotonated cis,cis,cis-1,2,3,4-cyclopentanetetracarboxylic acid ligands decompose at 420-750 °C to give the residue CoO in 1, ZnO + C in 2 and CoO + ZnO in 3. Complex 1 shows a complicated magnetic behavior with co-existence of antiferromagnetic exchange interactions between neighboring Co(2+) ions as well as strong spin-orbital coupling interactions for each Co(2+) ion; complex 3 exhibits a magnetically isolated high-spin Co(2+) ion behavior with strong spin-orbital coupling interactions.  相似文献   

3.
The complex [Co(bdt)(2)](-) (where bdt = 1,2-benzenedithiolate) is an active catalyst for the visible light driven reduction of protons from water when employed with Ru(bpy)(3)(2+) as the photosensitizer and ascorbic acid as the sacrificial electron donor. At pH 4.0, the system exhibits very high activity, achieving >2700 turnovers with respect to catalyst and an initial turnover rate of 880 mol H(2)/mol catalyst/h. The same complex is also an active electrocatalyst for proton reduction in 1:1 CH(3)CN/H(2)O in the presence of weak acids, with the onset of a catalytic wave at the reversible redox couple of -1.01 V vs Fc(+)/Fc. The cobalt-dithiolene complex [Co(bdt)(2)](-) thus represents a highly active catalyst for both the electrocatalytic and photocatalytic reduction of protons in aqueous solutions.  相似文献   

4.
Superoxide reduction by thiolate-ligated [FeII(SMe2N4(tren))]+ (1) involves two proton-dependent steps and a single peroxide intermediate, [FeIII(SMe2N4(tren))(OOH)]+ (2). An external proton donor is required, ruling out mechanisms involving H+ or H-atom abstraction from the ligand N-H. The initial protonation step affording 2 occurs with fairly basic proton donors (EtOH, MeOH, NH4+) in THF. More acidic proton donors are required to cleave the Fe-O(peroxide) bond in MeOH, and this occurs via a dissociative mechanism. Reaction rates are dependent on the pKa of the proton donor, and a common [FeIII(SMe2N4(tren))(MeOH)]2+ (3) intermediate is involved. Acetic acid releases H2O2 from 2 under pseudo-first-order conditions ([HOAc] = 138 mM, [2] = 0.49 mM) with a rate constant of 8.2 x 10(-4) s(-1) at -78 degrees C in MeOH. Reduction of 3 with Cp2Co regenerates the active catalyst 1.  相似文献   

5.
[Co(2,9-dimethyl-1,10-phenanthroline)(solvent)4]2+ ([Co(neo)]2+) undergoes a significant decrease in symmetry to form an inner-sphere surface complex when grafted directly on performed silica or introduced during the sol-gel process. The visible and X-ray absorption spectra of the surface adducts are interpreted in terms of a binding mode in which the Co(II) center has a highly distorted pseudo-C2v symmetry. The interaction of [Co(neo)]2+ with the silica surface was analyzed using an acid-base equilibrium relationship. Half-maximal surface binding was observed at pH ca. 6. Linear fits to the pH dependence data are consistent with inner-sphere binding of a single silanol group to the cobalt center. The formation of the surface species in tetramethoxyorthosilicate (TMOS) sol-gels required approximately 2 equiv of hydroxide anion per cobalt center, suggesting a two-proton-dependent binding event to form a species such as [Co(neo)(SiO)2]. Both sol-gel and silica samples showed essentially identical visible and X-ray absorption spectra, indicating formation of very similar surface adducts when the different synthesis procedures were employed. The maximal binding of [Co(neo)]2+ on three silica samples with different pore diameters and surface areas was compared. Increased binding was found to be inversely proportional to surface area and proportional to pore diameter, indicating a preference for less sterically demanding surface sites.  相似文献   

6.
Attempts to prepare heterobimetallic complexes in which 3d and uranium magnetic ions are associated by means of the Schiff bases H(2)L(i) derived from 2-hydroxybenzaldehyde or 2-hydroxy-3-methoxybenzaldehyde were unsuccessful because of ligand transfer reactions between [ML(i)] (M=Co, Ni, Cu) and UCl(4) that led to the mononuclear Schiff base complexes of uranium [UL(i)Cl(2)]. The crystal structure of [UL(3)Cl(2)(py)(2)] [L(3)=N,N'-bis(3-methoxysalicylidene)-ethylenediamine; py=pyridine] was determined. The hexadentate Schiff base ligand N,N'-bis(3-hydroxysalicylidene)-2,2-dimethyl-1,3-propanediamine (L) was useful for the synthesis of novel trinuclear complexes of the general formula [[ML(py)](2)U] (M=Co, Ni, Zn) or [[CuL(py)]M'[CuL]] (M'=U, Th, Zr) by reaction of [M(H(2)L)] with [M'(acac)(4)] (acac=MeCOCHCOMe). The crystal structures of the Co(2)U, Ni(2)U, Zn(2)U, Cu(2)U, and Cu(2)Th complexes show that the two ML fragments are orthogonal, being linked to the central actinide ion by the two pairs of oxygen atoms of the Schiff base ligand. In each compound, the UO(8) core exhibits the same dodecahedral geometry, and the three metals are linear. The magnetic study indicated that the two Cu(2+) ions are not coupled in the Cu(2)Zr and Cu(2)Th compounds. The magnetic behavior of the Co(2)U, Ni(2)U, and Cu(2)U complexes was compared with that of the Zn(2)U derivative, in which the paramagnetic 3d ion was replaced with the diamagnetic Zn(2+) ion. A weak antiferromagnetic coupling was observed between the Ni(2+) and the U(4+) ions, while a ferromagnetic interaction was revealed between the Cu(2+) and U(4+) ions.  相似文献   

7.
质子交换膜燃料电池(PEMFCs)电堆中阴极Pt基催化剂的高用量造成其成本居高不下,成为阻碍燃料电池汽车商业化推进的重要原因,因此开发低Pt、高活性的Pt基催化剂势在必行.Pt合金催化剂能够有效地降低Pt用量,并通过对合金颗粒的元素比例、晶面、粒径等实行精确调控,显著提升氧还原(ORR)催化活性.然而,目前常用的制备方法由于原料与制备成本高昂、过程复杂大都难以适应规模化生产需求.电化学方法通过控制施加的电流或电位控制晶体生长.在水体系中该方法已得到验证,但由于Pt化合物的热力学标准电极电位与过渡金属元素之间相差较大,且对于过渡金属来说,电负性大多小于铂,因此还原电位通常负于析氢电位,使得二者难以实现共沉积.有机体系中电位窗口比水体系大得多,Pt与电位较负的过渡金属可实现共沉积,采用小分子有机溶剂也可避免溶剂清洗问题,具有应用潜力.本文提出了一种简单的一步电沉积方法,选择易溶于水的N,N-二甲基甲酰胺(DMF)作为溶剂,将碳载体滴涂到玻碳电极上作为工作电极,通过电化学方法直接将Pt-Ni合金沉积到碳载体上,并利用物化表征与密度泛函理论(DFT)理论计算来探究共沉积机理.透射电镜表征结果表明,在不同的沉积电位下均可得到分散均匀、粒径适当的催化剂;且随着电位值降低,催化剂颗粒分散得更均匀,颗粒粒径不断减小.元素分布和晶面结果表明,铂镍元素均匀分布于颗粒中.所有样品均表现出优异的ORR性能,最高的面积比活性达到商业催化剂的6.85倍.将材料表征、电化学表征与DFT计算结合,建立起了铂镍合金生长过程的模型,并发现了有机体系中独特的成核-生长机理.将体系中的DMF换成超纯水,用同样的方法进行沉积,得到的催化剂颗粒团聚严重,说明DMF的使用能够避免颗粒团聚.在单独铂的体系中沉积发现,负载量极小,表明体系中镍前驱体的添加对于催化剂的沉积过程起到重要作用.电化学表征结果表明,在所选用的DMF有机体系中,镍的还原电位与铂的十分接近,但还原动力学更慢,趋向于先形成吸附原子后快速还原.由此可以推测,在二者合金的形成过程中,镍在碳载体表面的缓慢还原而形成的吸附原子能够成为铂还原的活性位点,从而降低了铂还原成核所需的能量,使得载体上的成核位点大大增加,这与DFT模拟结果一致.DFT建立了碳上镍的位点和铂的位点,分别在上面进行铂的还原,发现镍位点上比铂位点上更容易实现铂沉积.本文提出了铂镍共沉积的机理:在过电位(即还原能量)下,铂的还原动力学较镍稍快,于是铂先还原形成晶核,但难以达到生长的临界半径,于是单独铂体系中的沉积负载量很少.载体上还原的镍为铂还原提供了大量的活性位点,促进了铂还原,并与镍共沉积.Pt-Ni表面则进一步促进了铂的沉积和颗粒的生长.综上,本文提出了一种用于制备铂合金催化剂的有机电沉积体系,实现了单分散的碳载铂镍合金催化剂的一步制备.随后,本文将材料表征、电化学表征与DFT计算相结合,建立起了有机体系中铂镍合金成核-生长过程的机理模型.  相似文献   

8.
质子交换膜燃料电池(PEMFCs)电堆中阴极Pt基催化剂的高用量造成其成本居高不下,成为阻碍燃料电池汽车商业化推进的重要原因,因此开发低Pt、高活性的Pt基催化剂势在必行.Pt合金催化剂能够有效地降低Pt用量,并通过对合金颗粒的元素比例、晶面、粒径等实行精确调控,显著提升氧还原(ORR)催化活性.然而,目前常用的制备方法由于原料与制备成本高昂、过程复杂大都难以适应规模化生产需求.电化学方法通过控制施加的电流或电位控制晶体生长.在水体系中该方法已得到验证,但由于Pt化合物的热力学标准电极电位与过渡金属元素之间相差较大,且对于过渡金属来说,电负性大多小于铂,因此还原电位通常负于析氢电位,使得二者难以实现共沉积.有机体系中电位窗口比水体系大得多,Pt与电位较负的过渡金属可实现共沉积,采用小分子有机溶剂也可避免溶剂清洗问题,具有应用潜力.本文提出了一种简单的一步电沉积方法,选择易溶于水的N,N-二甲基甲酰胺(DMF)作为溶剂,将碳载体滴涂到玻碳电极上作为工作电极,通过电化学方法直接将Pt-Ni合金沉积到碳载体上,并利用物化表征与密度泛函理论(DFT)理论计算来探究共沉积机理.透射电镜表征结果表明,在不同的沉积电位下均可得到分散均匀、粒径适当的催化剂;且随着电位值降低,催化剂颗粒分散得更均匀,颗粒粒径不断减小.元素分布和晶面结果表明,铂镍元素均匀分布于颗粒中.所有样品均表现出优异的ORR性能,最高的面积比活性达到商业催化剂的6.85倍.将材料表征、电化学表征与DFT计算结合,建立起了铂镍合金生长过程的模型,并发现了有机体系中独特的成核-生长机理.将体系中的DMF换成超纯水,用同样的方法进行沉积,得到的催化剂颗粒团聚严重,说明DMF的使用能够避免颗粒团聚.在单独铂的体系中沉积发现,负载量极小,表明体系中镍前驱体的添加对于催化剂的沉积过程起到重要作用.电化学表征结果表明,在所选用的DMF有机体系中,镍的还原电位与铂的十分接近,但还原动力学更慢,趋向于先形成吸附原子后快速还原.由此可以推测,在二者合金的形成过程中,镍在碳载体表面的缓慢还原而形成的吸附原子能够成为铂还原的活性位点,从而降低了铂还原成核所需的能量,使得载体上的成核位点大大增加,这与DFT模拟结果一致.DFT建立了碳上镍的位点和铂的位点,分别在上面进行铂的还原,发现镍位点上比铂位点上更容易实现铂沉积.本文提出了铂镍共沉积的机理:在过电位(即还原能量)下,铂的还原动力学较镍稍快,于是铂先还原形成晶核,但难以达到生长的临界半径,于是单独铂体系中的沉积负载量很少.载体上还原的镍为铂还原提供了大量的活性位点,促进了铂还原,并与镍共沉积.Pt-Ni表面则进一步促进了铂的沉积和颗粒的生长.综上,本文提出了一种用于制备铂合金催化剂的有机电沉积体系,实现了单分散的碳载铂镍合金催化剂的一步制备.随后,本文将材料表征、电化学表征与DFT计算相结合,建立起了有机体系中铂镍合金成核-生长过程的机理模型.  相似文献   

9.
A new Zn(2+) fluorescent chemosensor N'-(3,5-di-tert-butylsalicylidene)-2-hydroxybenzoylhydrazine (H(3)L(1)) and its complexes [Zn(HL(1))C(2)H(5)OH](∞) (1) and [Cu(HL(1))(H(2)O)]CH(3)OH (2) have been synthesized and characterized in terms of their crystal structures, absorption and emission spectra. H(3)L(1) displays high selectivity for Zn(2+) over Na(+), K(+), Mg(2+), Ca(2+) and other transition metal ions in Tris-HCl buffer solution (pH = 7.13, EtOH-H(2)O = 8?:?2 v/v). To obtain insight into the relation between the structure and selectivity, a similar ligand 3,5-di-tert-butylsalicylidene benzoylhydrazine (H(2)L(2)), which lacks the hydroxyl group substituent in salicyloyl hydrazide compared with H(3)L(1), and its complex [Zn(2)(HL(2))(2)(CH(3)COO)(2)(C(2)H(5)OH)] (3), [Co(L(2))(2)][Co(DMF)(4)(C(2)H(5)OH)(H(2)O)] (4), [Fe(HL(2))(2)]Cl·2CH(3)OH (5), have also been investigated as a reference. H(3)L(1) exhibits improved selectivity for Zn(2+) compared to H(2)L(2). The findings indicate that the hydroxyl group substituent exerts an effect on the spectroscopic properties, complex structures and selectivity of the fluorescent sensor.  相似文献   

10.
将发展成熟的硅光伏电池与析氧催化剂(OEC)结合,在近中性环境下光解水制备了氢气。采用原位制备方法,在硅光伏材料上电沉积钴形成无定形Co-OEC膜,有效促进光生电荷分离,实现了利用AM1.5组合滤波片模拟太阳光照射下水分解制氢气。 结果表明,在K2B4O7缓冲溶液(pH=9.2)中,0 V(vs NHE)电压下沉积400 s形成的无定形Co-OEC具有良好的催化析氧效果,在无外加电压且模拟太阳光照射下,析氧电流密度可达1.7×10-2 A/cm2,产氧速率为1.28 mol/(h·m2),且具有良好的稳定性。  相似文献   

11.
The catalytic activity of carbon supported Pd-Co-Mo for the oxygen reduction reaction (ORR) in a single cell proton exchange membrane fuel cell (PEMFC) has been investigated at 60 degrees C and compared with data from commercial Pt catalyst and our previously reported Pd-Co-Au and Pd-Ti catalysts. The Pd-Co-Mo catalyst with a Pd:Co:Mo atomic ratio of 70:20:10 exhibits slightly higher catalytic activity like the Pd-Co-Au catalyst than the commercial Pt catalyst, but with excellent chemical stability unlike the Pd-Co-Au catalyst. The Pd-Co-Mo catalyst also exhibits better tolerance to methanol poisoning than Pt. Investigation of the catalytic activity of the Pd-Co-Mo system with varying composition and heat treatment temperature reveals that a Pd:Co:Mo atomic ratio of 70:20:10 with a heat treatment temperature of 500 degrees C exhibits the highest catalytic activity. Although the degree of alloying increases with increasing temperature from 500 to 900 degrees C as indicated by the X-ray diffraction data, the catalytic activity decreases due to an increase in particle size and a decrease in surface area.  相似文献   

12.
An electrochemical procedure of anodic deposition of cobalt oxyhydroxide film on a glassy carbon substrate in an alkaline medium (i.e. pH 11.6) is described. The electrodeposited film was obtained either by voltage cycling or by potentiostatic conditions using non-deaerated 0.1 M Na2CO3 solutions containing 40 mM tartrate ions and 4 mM CoCl2. The effects on the film formation and growth, such as tartrate–cobalt ratio, pH, applied potential, etc. were widely evaluated. The electrodeposition process, under anodic conditions and moderately alkaline solutions, most likely involves a redox transition Co(II)→Co(III)/Co(IV) with destruction of the tartrate complex and formation of insoluble oxide/hydroxide cobalt species on the glassy carbon surface. The resulting cobalt oxyhydroxide films were characterised by cyclic voltammetry (CV) in 0.1 M NaOH solutions and by scanning electron microscopy (SEM) analysis after different strategies of preparation and various electrochemical treatments. The electrochemical activity of the deposited films was checked using various organic molecules as model compounds.  相似文献   

13.
Effect of electrodeposition conditions on the nucleation of gold crystals at a glassy-carbon cathode in an NaCl–KCl–CsCl melt at 500–700°C is studied by a galvanostatic technique. The maximum nucleation overvoltage can reach 550 mV. At low current densities, the surface of glassy carbon undergoes activation due to spontaneous deposition of gold crystals. The overvoltage vs. time curve for a melt containing an elevated (as compared with equilibrium) amount of chlorine exhibits additional wave that precedes the nucleation peak. The wave is probably connected with the recharge of ions of Au(III) to Au(I). The oxide film disappears from the cathode surface, thus facilitating the nucleation process. On a modeling electrode with a current leakage, an increase in the maximum overvoltage is discovered, which is not connected with the cathode activation. Exchange currents for the discharge-ionization of gold at the growing nuclei are estimated.  相似文献   

14.
A cobalt bis(iminopyridine) complex is a highly active electrocatalyst for water reduction, with an estimated apparent second order rate constant k(app) ≤ 10(7) M(-1)s(-1) over a range of buffer/salt concentrations. Scan rate dependence data are consistent with freely diffusing electroactive species over pH 4-9 at room temperature for each of two catalytic reduction events, one of which is believed to be ligand based. Faradaic H(2) yields up to 87 ± 10% measured in constant potential electrolyses (-1.4 V vs SCE) confirm high reactivity and high fidelity in a catalyst supported by the noninnocent bis(iminopyridine) ligand. A mechanism involving initial reduction of Co(2+) and subsequent protonation is proposed.  相似文献   

15.
In methanol or chloroform/methanol solutions, reactions of Cltpy or MeOtpy (Rtpy = 4'-R-2,2':6',2'-terpyridine) with CoX(2)·xH(2)O (X(-) = Cl(-), [OAc](-), [NO(3)](-) or [BF(4)](-)) result in the formation of equilibrium mixtures of [Co(Rtpy)(2)](2+) and [Co(Rtpy)X(2)]. A study of the solution speciation has been carried out using (1)H NMR spectroscopy, aided by the dispersion of signals in the paramagnetically shifted spectra; on going from a low- to high-spin cobalt(II) complex, proton H(6) of the tpy ligand undergoes a significant shift to higher frequency. For R = Cl and X(-) = [OAc](-), increasing the amount of CD(3)OD in the CD(3)OD/CDCl(3) solvent mixture affects both the relative proportions of [Co(Cltpy)(2)](2+) and [Co(Cltpy)(OAc)(2)] and the chemical shifts of the (1)H NMR resonances arising from [Co(Cltpy)(OAc)(2)]. When the solvent is essentially CDCl(3), the favoured species is [Co(Cltpy)(OAc)(2)]. For the 4'-methoxy-2,2':6',2'-terpyridine, the speciation of mono- and bis(terpyridine)cobalt(II) complexes depends upon the anion, solvent and ligand:Co(2+) ion ratio. The (1)H NMR spectrum of [Co(MeOtpy)(2)](2+) is virtually independent of anion and solvent. In contrast, the signals arising from [Co(MeOtpy)X(2)] depend on the anion and solvent. In the case of X(-) = [BF(4)](-), we propose that the mono(tpy) complex formed in solution is [Co(MeOtpy)L(n)](2+) (L = H(2)O or solvent, n = 1-3). The formation of mono(tpy) species has been confirmed by the solid state structures of [Co(Cltpy)(OAc-O)(OAc-O,O')], [Co(MeOtpy)(OAc-O)(OAc-O,O')], [Co(MeOtpy)(NO(3)-O)(2)(OH(2))] and [Co(MeOtpy)Cl(2)]. The single crystal structure of the cobalt(III) complex [Co(Cltpy)Cl(3)]·CHCl(3) is also reported.  相似文献   

16.
A series of CoO(x)-doped silica xerogels with various Co(2+) loadings (Co/Si = 0, 1, 2, 4, 6, and 10 mol %) has been prepared. All xerogels exhibit large (800-1050 m(2)/g) surface areas. Narrow pore size distributions with pore size maxima around 3 nm are characteristic for Co/Si = 1, 2, 4, 6, 10 samples. As-prepared CoO(x)/SiO(2) xerogels show high catalytic activity in the air oxidation of gaseous acetaldehyde at room temperature. Carbon dioxide and trace amounts of methane are the only products detected in the gas phase. Acetic acid, a less volatile product, resides on the surface of the xerogels but can slowly desorb. The formation of CO(2) begins after an induction period. The beginning of CO(2) production coincides with the conversion of Co(2+) incorporated in the SiO(2) framework into Co(3+). Thermogravimetry/gas chromatography/mass spectrometry analysis, UV-vis and FTIR spectroscopies, as well as kinetic measurements are employed for CoO(x)/SiO(2) catalyst characterization. A possible mechanism of the reaction is discussed.  相似文献   

17.
The S-bonded sulfenamide isomers have been prepared by the known reaction of hydroxylamine-O-sulfonate with (en)(2)Co(III) thiolate complexes of aminoethanethiol, cysteine, and penicillamine and the cis dithiolate formed by N,N'-ethylene-di-penicillamine (EDP) and Co(III). It is shown that the sulfenamides undergo linkage isomerization in alkaline solution to produce their respective N-bonded linkage isomers. The addition of acid yields the protonated N-bonded isomers. The structures of [(en)(2)Co(NH(2)S(CH(2))(2)NH(2))](2)(S(2)O(6))(3) and [Co((NH(2)S)(2)EDP)]Br have been determined by X-ray crystallography, and the pK(a) of [(en)(2)Co(NH(2)S(CH(2))(2)NH(2))](3+) has been determined by spectrophotometry. The pH dependencies of the kinetics of the linkage isomerization reactions have been studied and yield pK(a) values of the S-bonded isomers. The (en)(2)Co systems give only the acid-stable N,N' isomer at equilibrium, whereas the EDP complex gives a mixture of N,N' and N,S isomers at pH 7-9.  相似文献   

18.
The present work demonstrates the possibilities and the limits of the in situ electrochemical scanning tunneling microscopy for investigation of nucleation processes in magnetic fields on the examples of Cu and Co electrodeposition onto Au(111) electrodes from sulfate electrolytes with pH 3. Cyclic voltammograms of Cu in the underpotential range (UPD) exhibit no significant change in the cathodic and anodic peaks recorded in magnetic fields parallel to the surface. In magnetic fields of a permanent magnet, the reconstruction of Au has been annihilated during UPD of Cu. In the overpotential range, the dissolution of Cu is inhibited. This triggers the formation of a Cu–Au surface alloy. The UPD deposition of Co onto Au(111) could be proven without magnetic field, which leads to the formation of two monolayers. The nucleation in an applied field could not be observed due to higher induced fluctuations and microconvective effects. Contribution to special issue “Magnetic field effects in Electrochemistry”.  相似文献   

19.
A 2,4'-bithiazole group has been covalently attached to the Co(III) complex of a designed ligand PMAH that mimics the metal-binding locus of the antitumor drug bleomycin (BLM). The deprotonated PMA(-) ligand binds Co(III) via five nitrogens located in primary and secondary amines, a pyrimidine and an imidazole ring, and a peptide moiety. The 2,4'-bithiazole group is tethered to the [Co(PMA)](2+) unit via an imidazole that is connected to the bithiazole moiety with a (CH(2))(3) spacer. The structure of this hybrid analogue, namely, [Co(PMA)(Bit)]Cl(2) (7, Bit = 2'-methyl-2,4'-bithiazole-4-carboxamido-N'-(3-propyl)imidazole) has been established by spectroscopic techniques. 7 promotes photocleavage of DNA at micromolar concentrations. Unlike simpler analogues like [Co(PMA)(H(2)O)](2+) and [Co(PMA)Cl)](+) which induce random DNA cleavage upon UV illumination, 7 exhibits sequence specificity in the DNA photocleavage reaction. Intriguing is to note that 7 exhibits the same 5'GG-N3' sequence preference as another hybrid analogue [Co(PMA)(Int-A)]Cl(2) (6, Int-A = acridine-9-carboxamido-N'-(3-propyl)imidazole) that contains an acridine moiety as the DNA-binding group. The observed sequence specificity of 6 and 7 therefore does not reflect the sequence preferences of the DNA-binding groups (acridine and bithiazole). The results indicate that the metalated core of the hybrid analogues, i.e., the [Co(PMA)](2+) unit is the key factor in determining their sequence specificity.  相似文献   

20.
Co(II) solution species containing 1 equiv of phenanthroline (phen), 2-methyl-1,10-phenanthroline (MMP), or 2,9-dimethyl-1,10-phenanthroline (DMP) ligand formed inner-sphere surface complexes when grafted on silica. The speciation on the silica surface depended on both the pH of the grafting solution and the steric bulk of the ligand. [Co(DMP)](2+) formed tetrahedral surface adducts exclusively, with a 1:1 ligand-Co ratio. These surface adducts were first detectable at pH values above 5.1. [Co(MMP)](2+) and [Co(phen)](2+) formed exclusively octahedral adducts on the surface with a 1:1 ligand-Co ratio at pH values below 5. The [Co(MMP)](2+) complex formed a tetrahedral adduct initially at pH 6 and increasingly as the pH was raised. The [Co(phen)](2+) complex did not produce a comparable tetrahedral surface species under any conditions. Instead, mixtures of octahedral surface species with both 1:1 and 2:1 ligand-Co ratios began to form at pH values above 6. Taken together, the results indicated that the development of tetrahedral stereochemistry was strongly influenced by steric factors in the presence of a nitrogen-donating ligand. All three phenanthroline derivatives promoted surface binding of the Co(II) ion adducts, so that maximal binding occurred at lower pH values than for binding of [Co(H(2)O)(6)](2+), which formed exclusively tetrahedral adducts.  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