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 218 毫秒
1.
The structure of rhombohedral (R) iron(III) tris­[di­hydrogen­phosphate(I)] or iron(III) hypophosphite, Fe(H2PO2)3, has been determined by single‐crystal X‐ray diffraction. The structure consists of [001] chains of Fe3+ cations in octa­hedral sites with symmetry bridged by bidentate hypophosphite anions.  相似文献   

2.
Anodic oxidation of sodium hypophosphite on smooth Pd and Pd/Pt electrodes and a Pd membrane is studied. Thei vs.E curves for the Pd electrode exhibit two anodic current peaks. One is caused by oxidation of H2PO 2 - , and the other, by simultaneous ionization of Pd and oxidation of H2PO 2 - . The hypophosphite ion adsorbed on the Pd surface hinders the formation of the passive film. This brings about a rapid dissolution of Pd in the oxygen region and its subsequent deposition with the formation of palladium black. The oxidation probably includes a slow heterogeneous chemical reaction, specifically, a cleavage of the P-H bond of the hypophosphite ion. The change in the reaction stoichiometry following an increase in solution pH and in anodic polarization is probably due to changing conditions of the H2PO 2 - adsorption and the number of adsorption sites occupied by H2PO 2 - on the surface. Following an increase in polarization, the phosphite ion may undergo oxidation to phosphate. Deceased.  相似文献   

3.
The redox catalysis of oxygen reduction was performed on a platinum rotating disk electrode. The Fe(III)/Fe(II)/H2SO4 system at different pH's was used as a MEDIATOR. The catalytic effect of mediator was directly related to the solvation sphere of Fe(III) and Fe(II). Only the redox couple FeHSO 4 2+ /FeHSO 4 + (pH<0) showed a catalytic effect on oxygen reduction.  相似文献   

4.
Sulfite reductase (SiR) catalyzes a six electron and six proton reduction of sulfite to sulfide. Similarly to the cytochrome P450 (cytP450) family, the active site in SiR contains a (partially reduced) heme bound axially to a cysteinate ligand—though with an extra Fe4S4 cluster. Fe(III) SO2−, Fe(III) SOH, and Fe(III) SO(H2) intermediates have been proposed for the catalytic cycle of SiR, leading to a formally Fe(V)S species—akin to the widely accepted reaction mechanism in cytP450. Here, density functional theory (DFT) data is reported for of such FeSO(H2) intermediates. The Fe(III) SO2− models display relatively high energies for homolytic bond breaking compared to their isomeric oxygen‐bound Fe(III) OS2− models, and thus offer a better alternative in terms of avoiding radical side products able to induce enzyme suicide. This could be due to the fact that the (iron‐bound) sulfur is more active from a redox standpoint compared to oxygen, thus permitting the departing oxygen to maintain a redox‐inert state. Di‐protonation of the oxygen is computed to lead to a compound I type Fe(IV)S coupled to a porphyrin radical anion—consistent with an intermediate previously observed by x‐ray crystallography.  相似文献   

5.
The syntheses and structures of four new compounds are described. Two of these compounds are the anhydrous and dihydrate chloride salts of the diamagnetic bis(2,6-diacetylpyridinedioxime)iron(II) cation, [Fe(DAPDH2)2]2+. In this complex cation the DAPDH2 ligand binds to the iron, as expected, through its three nitrogen atoms leaving the four oxime oxygen atoms protonated and uncoordinated. The third compound is (AsPh4)2[Fe2OCl6], a new salt of the well-known oxo-bridged diiron complex, [Fe2OCl6]2?. The synthesis of (AsPh4)2[Fe2OCl6] is a high yield, straightforward, one-step preparation starting with AsPh4Cl and ferrous chloride in methanol. In this synthesis Fe(II) is oxidized to Fe(III) by atmospheric O2. The fourth new compound is the novel and unexpected triiron complex [Fe(DAPDH)2Fe2OCl4]. This complex is derived from [Fe(DAPDH2)2)]2+ and [Fe2OCl6]2? by removing the H+ from each of two adjacent oxime oxygen atoms of the former and one Cl? from each of the Fe(III) ions of the latter. The resulting neutral fragments, Fe(DAPDH)2 and Fe2OCl4, are joined via bonds linking the two oxime oxygen atoms to the two Fe(III) ions giving rise to an unusual eight membered chelate ring containing three iron ions, two nitrogen atoms and three oxygen atoms, one of which is the bridge between the two Fe(III) ions.  相似文献   

6.
The phosphoryl donor ligand (tBuNH)3PO (PO) was used for preparation of new tin(IV), Cl2Ph2Sn(PO)2 (1), and iron(III), Fe(PO)2(NO3)3 (2), complexes. These complexes are the first examples of using a phosphoric triamide containing a secondary nitrogen atom, [RNH]3P(O), for preparation of an organotin(IV) complex of the type ([RNH]3P(O))2X2Ph2Sn, X = halide, and an iron(III) complex. In 1, the Sn coordination geometry is octahedral with the pair of similar ligands in a trans orientation. The Fe center in 2 is seven-coordinated with the two phosphoramide ligands in a trans fashion, too. This article also reviews the structures of analogous complexes with phosphoric triamide ligands, deposited in the CSD, aiming to classify hydrogen bond patterns in this category of compounds. Moreover, it is tried to find a relationship between the H-bond patterns in complexes and the related free ligands.  相似文献   

7.
In a stirred batch reaction, Fe(phen)32+ ion behaves differently from Ce(III) or Mn(II) ion in catalyzing the bromate‐driven oscillating reaction with ethyl hydrogen malonate [CH2COOHCOOEt, ethyl hydrogen malonate (EHM)]. The effects of N2 atmosphere, concentrations of bromate ion, EHM, metal ion catalyst, sulfuric acid, and additive (bromide ion or bromomalonic acid) on the pattern of oscillations were investigated. The kinetic study of the reaction of EHM with Ce(IV), Mn(III), or Fe(phen)33+ ion indicates that under aerobic or anaerobic conditions the order of reactivity toward reacting with EHM is Mn(III) > Ce(IV) ≫ Fe(phen)33+, which follows the same trend as that of the malonic acid system. The presence of the ester group in EHM lowers the reactivity of the two methylene hydrogen atoms toward bromination or oxidation by Ce(IV), Mn(III), or Fe(phen)33+ ion. No good oscillations were observed for the BrO3−‐CH2(COOEt)2 reaction catalyzed by Ce(III), Mn(II), or Fe(phen)32+ ion. A discussion of the effects of oxygen on the reactions of malonic acid and its derivatives (RCHCOOHCOOR′) with Ce(IV), Mn(III), or Fe(phen)33+ ion is also presented. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 52–61, 2000  相似文献   

8.
Bis(trimethylsilyl)hypophosphite und Alkoxycarbonylphosphonous Acid Bis(trimethylsilyl) esters as Building Blocks in Organophosphorus Chemistry The oxidation of pure bis(trimethylsilyl)hypophosphite ( BTH ) with chalcogenides forming (Me3SiO)2P(X)H (X = O, S, Se, Te) is described as well as its reactions with alkylhalides RX (X = Cl, Br, I) and Cl? C(O)OR (R = Me, Et, Bzl). By reaction with oxygen, sulfur, and selenium the alkoxycarbonylphosphonous acid bis(trimethylsilyl)esters form RO? C(O)? P(X)(OSiMe3)2 (X = O, S, Se) whereas with Cl? C(O)OR the bis(alkoxycarbonyl)-phosphinic acid trimethylsilylesters are obtained. After partial hydrolysis the resulting instable RO? C(O)? P(O)H(OSiMe3) gives RO? C(O)? P(O)(OSiMe3)? CH2? NH? A? COOR′ (A = CH2, CH2CH2, CHCH3, CH2CH2SH, CHCH(CH3)2,…) when allowed to react with hexahydro-s-triazines of the aminoacid esters. Reactions of the alkoxycarbonyl-P-silylesters with NaOR or NaOH result in the corresponding mono-, di-, or trisodium salts. With mineral acids decarboxylation occurs, but H? P(O)(OH)? CH2? NH? A? COOH can be obtained, too. The structure of the compounds described are discussed by their n.m.r. data.  相似文献   

9.
In this study, the role of oxygen in the regeneration of Fe(III) during the degradation of atrazine in UV/Fe(III) process was studied. The degradations of atrazine in UV/Fe(III) and UV-photolysis processes in the presence and absence of oxygen were compared. The results showed that the degradations of atrazine in these processes followed the pseudo-first-order kinetics well. The process exhibiting the highest rate constant (k) was UV/Fe(III)/air process, because k-value for UV/Fe(III)/air process was about 1.47, 2.23 and 2.56 times of those for UV/Fe(III)/N2, UV/air and UV/N2 processes, respectively. The degradation of atrazine was enhanced by oxygen in UV/Fe(III) process and the enhancement was more remarkable at higher initial concentrations of Fe(III). The investigation into the changes of Fe(III) concentrations demonstrated that the presence of oxygen led to the regeneration of Fe(III), which resulted in the enhancement of atrazine degradation. With air bubbling, the ferric ions were 25% more than those with N2 bubbling. The experimental data showed the regeneration of Fe(III) required the excited organic molecules and oxygen and on the basis of these results, the regeneration mechanism of Fe(III) was proposed. It was also found that due to the oxidation of Fe(II), the degradation of atrazine in UV/Fe(II)/air process was effective at a low Fe(II) concentration of 7 mg/L, similar to that in UV/Fe(III)/air process. This study makes clear the role of oxygen in the regeneration of Fe(III), and thus it provides a guide to reduce the input of Fe(III) and is helpful to the application of UV/Fe(III) process in practice.  相似文献   

10.
The action of ozone on a suspension of Fe(III) hydroxide in alkaline solutions was studied by the spectrophotometric method. A partial conversion of Fe(III) to Fe(VI) is observed at Fe(III) concentrations exceeding 2 mmol l−1. The tenfold increase of the initial Fe(III) concentration raises the Fe(VI) yield by a factor of 2–3. The mechanism of the process includes the decomposition of ozone with the formation of ozonide ions, which oxidize Fe(III) up to Fe(IV), Fe(V), and Fe(VI) in parallel with their conversion to O2 and HO2. Fe(VIII) is not formed.  相似文献   

11.
The reaction between Fe(III) and dopamine in aqueous solution in the presence of Na2S2O3 was followed through UV–Vis spectroscopy, pH and oxy-reduction potential (Eh) measurements. The formation and quick disappearing of the complex [Fe(III)HL1−]2+, HL1− = monoprotonated dopamine was observed with or without S2O3 2− at pH 3. An unexpected reaction occurs in presence of thiosulfate forming the stable anion complex [Fe(III)(L2−)2]1−, L2− = dopacatecholate (λ = 580 nm) and the auto-increasing of the pH, from 3 to 7. It was proposed that H+ and molecular oxygen are consumed by free radical thiosulfate formed during the reaction.  相似文献   

12.
Salicylaldehyde or 5-bromosalicylaldehyde react with 2,3-diaminophenol to give two unsymmetrical Schiff-bases H2L1, H2L2, respectively. With Fe(III) and Co(II), these ligands lead to four complexes: Fe(III)ClL1, Fe(III)ClL2, Co(II)L1, Co(II)L2. The structures of these complexes were determined by mass spectroscopy, infrared and electronic spectra. Cyclic voltammetry in dimethylformamide (DMF) showed irreversible waves for both ligands. In the same experimental conditions, Fe(III)ClL1 exhibited a reversible redox couple Fe(III)/Fe(II) while the three other complexes showed quasi-reversible systems. The behavior of some of these complexes in the presence of dioxygen and the comparison with cytochrome P450 are described.  相似文献   

13.
The preparation by hydrothermal reaction and the crystal structure of the iron(III) carboxyethylphosphonate of formula [NH4][Fe2(OH){O3P(CH2)2CO2}2] is reported. The green-yellow compound crystallizes in the monoclinic system, space group Pc(n.7), with the following unit-cell parameters: a=7.193(3) Å, b=9.776(3) Å, c=10.17(4) Å and β=94.3(2)°. It shows a typical layered hybrid organic-inorganic structure featuring an alternation of organic and inorganic layers along the a-axis of the unit cell. The bifunctional ligand [O3P(CH2)2CO2]3− is deprotonated and acts as a linker between adjacent inorganic layers, to form pillars along the a-axis. The inorganic layers are made up of dinuclear Fe(III) units, formed by coordination of the metal ions with the oxygen atoms originating from the [O3P−]2− end of the carboxyethylphosphonate molecules, the oxygen atoms of the [−CO2] end group of a ligand belonging to the adjacent layer and the oxygen atom of the bridged OH group. Each Fe(III) ion is six-coordinated in a very distorted octahedral environment. Within the dimer the Fe-Fe separation is found to be 3.5 Å, and the angle inside the [Fe(1)-O(11)-Fe(2)] dimers is ∼124°. The resulting 3D framework contains micropores delimited by four adjacent dimers in the (bc) planes of the unit cell. These holes develop along the a-direction as tunnel-like pores and [NH4]+ cations are located there. The presence of the μ-hydroxo-bridged [Fe(1)-O(11)-Fe(2)] dimers in the lattice is also responsible for the magnetic behavior of the compound at low temperatures. The compound contains Fe3+ ions in the high-spin state and the two Fe(III) ions are antiferromagnetic coupled. The J/k value of −16.3 K is similar to those found for other μ-hydroxo-bridged Fe(III) dimeric systems having the same geometry.  相似文献   

14.
The electrochemical electron-transfer rate constants for the redox systems Fe(IV)L3+/Fe(III)L3 (L=N,N-disubstituted dithicarbamate ion) and Fe(III)L3/Fe(II)L3? with a variety of substituents were measured at a platinum electrode in acetonitrile with the galvanostatic double-pulse method. It is known that each of the Fe(III) complexes exists both in a highspin state 6A1 and a low-spin state 2T2 in equilibirium of which position is widely changed by a subtle change in substituent. The standard rate constants for Fe(IV)L3+/Fe(III)L3 were larger or smaller than those for Fe(III)L3/Fe(II)L3? according as the Fe(III)L3 complexes are predominantly low- or high-spin complexes. Since the Fe(IV) and Fe(II) complexes are low-and high-spin complexes respectively, these findings suggest that electrochemical electron-transfer reactions accompanied by a spin-state change are slower than those without it. Such spin-state effect on electrode reactions has rarely been discussed so far.  相似文献   

15.
Cyclic voltammetry, electrochemical impedance spectroscopy, and rotating disk electrode voltammetry have been used to study the effect of chloride ions on the dissolved oxygen reduction reaction (ORR) on Q235 carbon steel electrode in a 0.02 M calcium hydroxide (Ca(OH)2) solutions imitating the liquid phase in concrete pores. The results indicate that the cathodic process on Q235 carbon steel electrode in oxygen-saturated 0.02 M Ca(OH)2 with different concentrations of chloride ions contain three reactions except hydrogen evolution: dissolved oxygen reduction, the reduction of Fe(III) to Fe(II), and then the reduction of Fe(II) to Fe. The peak potential of ORR shifts to the positive direction as the chloride ion concentration increases. The oxygen molecule adsorption can be inhibited by the chloride ion adsorption, and the rate of ORR decreases as the concentration of chloride ions increases. The mechanism of ORR is changed from 2e and 4e reactions, occurring simultaneously, to quietly 4e reaction with the increasing chloride ion concentration.  相似文献   

16.
Energy surfaces of low-lying states of planar complexes [Fe(C2H2X2)2]n with X=O and S and total charges n varying between + 3 and − 2 have been investigated by quantum chemical ab initio MO-SCF calculations of double zeta quality. Through studies of such properties as geometry, electron density distribution and molecular orbital energies it has been concluded that some of the states can be referred to as Fe (III) and others as Fe (II) states. The lowest states are found to be those with n = − 1 and 0 for X = O and n = − 1 and − 2 for X = S. For both the oxygen and the sulfur complexes Fe (III) and Fe (II) states are very close in energy. Supposing other electron acceptors and donors to be available, possible schemes for redox reactions between metal and ligand are suggested.  相似文献   

17.
In this study, an environmentally friendly complexing agent, S,S′-ethylenediamine-N,N′-disuccinic acid (EDDS), was applied in Fe(III)-mediated activation of persulfate (PS), and the degradation performance of trichloroethylene (TCE) was investigated. The effects of PS concentration, Fe(III)/EDDS molar ratio, and inorganic anions on TCE degradation were evaluated, and the generated reactive oxygen species responsible for TCE removal were identified. The results showed that nearly complete TCE degradation was achieved with PS of 15.0 mM and a molar ratio of Fe(III)/EDDS of 4:1. An increase in PS concentration or Fe(III)/EDDS molar ratio to a certain value resulted in enhanced TCE degradation. All of the anions (Cl?, HCO3 ?, SO4 2?, and NO 3 ? ) at tested concentrations had negative effects on TCE removal. In addition, investigations using radical probe compounds and radical scavengers revealed that sulfate radicals (SO 4 ·? ), hydroxyl radicals (·OH), and superoxide radical anions (O 2 ·? ) were all generated in the Fe(III)–EDDS/PS system, and ·OH was the primary radical responsible for TCE degradation. In conclusion, the Fe(III)–EDDS-activated PS process is a promising technique for TCE-contaminated groundwater remediation.  相似文献   

18.
《Electroanalysis》2003,15(2):97-102
The oxidation of dimethyl sulfoxide (DMSO) to dimethyl sulfone (DMSO2) is representative of numerous anodic oxygen‐transfer reactions of organosulfur compounds that suffer from slow kinetics at noble metal electrodes. Anodic voltammetric data for DMSO are examined at various RuO2‐film electrodes prepared by thermal deposition on titanium substrates. The response for DMSO is slightly larger at RuO2 films prepared in a flame as compared with films prepared in a furnace; however, temperature is more easily controlled in the furnace. Doping of the RuO2 films with Fe(III) further improves the sensitivity of anodic response for DMSO. Optimal response is obtained at an Fe(III)‐doped RuO2‐film electrode prepared using a deposition solution of 50 mM RuCl3 and 10 mM FeCl3 in a 1 : 1 mixture of isopropanol and 12 M HCl at an annealing temperature of 450 °C. The Levich plot (i vs. ω1/2) and Koutecky‐Levich plot (1/i vs. 1/ω1/2) of amperometric data for the oxidation of DMSO at an Fe(III)‐doped RuO2‐film electrode configured as a rotated disk are consistent with an anodic response controlled by mass‐transport processes at low rotational velocities. Flow injection data demonstrate that Fe(III)‐doped RuO2‐film electrodes exhibit detection capability for methionine and cysteine in addition to DMSO. Detection limits for 100‐μL injections of the three compounds are ca. 3.2×10?4 mM, i.e., ca. 32 pmol.  相似文献   

19.
A Debye temperature θD of 378 (±5K) has been obtained by applying a simplified Debye model to the57Fe Mössbauer spectra of 60CaO·39Ga2O3·57Fe2O3 glass. The θD value is comparable to those (280–580 K) obtained so far in several oxide glasses, glass-ceramics, and ceramics in which Mössbauer atoms are covalently bonded to oxygen atoms and play a role of network former. It proves that Fe(III) atoms occupy the substitutional sites of Ga(III) constituting distorted GaO4 tetrahedra.  相似文献   

20.
The behavior and conditions of liquid‐liquid extraction‐separation of Fe(III) by ammonium thiocyanate‐H2O‐n‐propyl alcohol system in the presence of NaCl were studied, and the possible reactive mechanism of extraction of Fe(III) was deduced. The study showed that, in the presence of a given amount of NaCl, phases were separated thoroughly between n‐propyl alcohol and water. In the process of phase separation, the complex [Fe(SCN)n](3‐n) formed by NH4SCN and Fe(III) was quantitatively extracted into the n‐propyl alcohol phase. The extracted Fe(III) exists in the n‐propyl alcohol phase mainly as the forms of Fe(SCN)2+ and Fe(SCN)3. Also, the relationship between extraction yield of Fe(III) and the amount of NH4SCN agreed well with the quadratic equation E = 0.54 + 58.14x ? 8.39x2 (E and x represent the recovery rate of Fe(III) and the volume (mL) of 0.1 M NH4SCN respectively). The quadratic R‐Square is 0.9990. With this method, Fe(III) can be completely separated from Co(II), Ni(II), Mn(II), Al(III), Bi(III) and Cd(II) at pH 1.0?2.0. The present method was applied in determining Fe(III) in samples with satisfactory results such as relative standard deviation from 2.06% to 2.89% and recovery rate in the range of 98.4?101.4%.  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