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 15 毫秒
1.
The isotope exchange reactions of malonic acid and a malonate ion were investigated in acidic and basic D2O solutions, respectively, using 1H NMR spectroscopy. The isotope exchange reaction of malonic acid is inhibited by the presence of DNO3 (0–3 M) and DSO4? ion (0–0.1 M), whereas it is catalyzed by the presence of DSO4? ion (> 0.2 M), D3PO4, D2PO4? ion or DPO42– ion. The order of relative reactivity for catalyzing the isotope reaction of malonic acid in D2O is DPO42– > D2PO4? > D3PO4 > DSO4? > DNO3. The rate of the isotope exchange reaction of malonate ion in D2O decreases to a minimum and then increases with increased [NaOD]0. The mechanism of the isotope exchange reaction of malonic acid in acidic D2O is different from the general acid-catalyzed mechanism generally observed for organic acids like acetic and dichloroacetic acids. The bimalonate ion plays an important role in the isotope exchange reactions of this system.  相似文献   

2.
3.
The acetonylgold(III) compound [Au(ppy)(CH2COCH3)Cl] (1) (ppy = 2-phenylpyridine) was unexpectedly obtained during the crystallization process of Au(III) lactate complex [Au(ppy)(CH3CHOHCOO)Cl]. This new structure prompted us to further study the role of Au(III) complexes on the carbon-hydrogen activation of ketones. Complex [Au(ppy)(CH2COCH3)NO3] (2) was synthesized by reacting [Au(ppy)(NO3)2] with acetone while the ketonyl Au(III) complex [Au(apd)Cl2] (3) (Hapd = 2-acetylpyridine) was obtained through carbon-hydrogen bond activation of the acetyl group. The crystal structures of 1 and 2 have common features: a square-planar Au(III) centre coordinated by one five-membered chelate ring, one acetonyl ligand and one anion (chloride or nitrate). Both structures show that carbon-hydrogen activation of acetone by 2-phenylpyridine-Au(III) complexes leads to the formation of acetonyl-Au(III) complexes. The Au-CH2 bond lengths (2.067(7) Å, 1 and 2.059(5) Å, 2) are similar to each other but longer than the Au-C (phenyl) bond lengths. The two softest ligands (carbanion) are also cis to each other in the thermodynamically most stable isomer. In complex 3, the σ-bonded acetyl group is confirmed by 13C DEPT NMR spectroscopy.  相似文献   

4.
A cost‐effective and sensitive colorimetric method was described for the determination of chromium(III) ion (Cr3+) by using ethylenediaminetetraacetic acid functionalized gold nanoparticles (EDTA‐AuNPs) as a probe. The stable and dispersed EDTA‐AuNPs were prepared by reducing HAuCl4 with sodium borohydride in presence of EDTA as a capping agent. Upon the addition of Cr3+, the colour of EDTA‐AuNPs solution changed from red to violet, which was in response to the surface plasmon absorption of dispersed and aggregated EDTA‐AuNPs. The procedure allowed the determination of Cr3+ in the range of 0.1–1.0 mol/L. The limit of detection for Cr3+ was 0.08 mol/L. The relative standard deviation was 2.5 % for eight repeated measurements of 0.6 mol/L Cr3+ solution. The method was applied to the determination of Cr3+ in water samples.  相似文献   

5.
A samarium(III) complex (SmAOA) of 5‐aminoorotic acid (HAOA) was synthesized by reaction of the respective inorganic salt and ligand in a molar ratio of 1:3. The structure of the complex was determined by means of elemental analysis and UV–visible, NMR, Fourier transform infrared (FT‐IR) and FT Raman spectroscopies. Significant differences in the FT‐IR and FT Raman spectra of the complex were observed as compared to the spectra of the free ligand. Detailed vibrational analysis of HAOA and Sm(III)–AOA systems revealed that the binding mode in the complex was bidentate through the carboxylic oxygen atoms. The newly synthesized SmAOA showed antioxidant properties. The antioxidant activities of both HAOA and SmAOA were related to their electron donor properties. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

6.
Six novel organometallic half sandwich complexes [(η5‐C5Me5)M(L1–3)Cl]Cl.2H2O were synthesized using [{(η5‐C5Me5)M(μ‐Cl)Cl2], where M = Ir (III)/Rh (III) and L1–3 = three pyridyl pyrimidine based ligands; and characterized by NMR, Infra‐red spectroscopy, conductance, elemental and thermal analysis. The complex‐DNA binding mode and/or strength evaluated using absorption titration, electrochemical studies and hydrodynamic measurement proposed intercalative binding mode, which was also confirmed by molecular docking study. Differential pulse voltammetry and cyclic voltammetry studies indicated an alteration in oxidation and reduction potentials of complexes (M+4/M+3) in presence of CT‐DNA. The metal complexes can cleave plasmid DNA as proposed in gel electrophoretic analysis. The LC50 values of complexes evaluated on brine shrimp suggested their potent cytotoxic nature.  相似文献   

7.
In a stirred batch experiment and under aerobic conditions, ferroin (Fe(phen)32+) behaves differently from Ce(III) or Mn(II) ion as a catalyst for the Belousov‐Zhabotinsky (BZ) reaction with allylmalonic acid (AMA). The effects of bromate ion, AMA, metal‐ion catalyst, and sulfuric acid on the oscillating pattern were investigated. The kinetics of the reaction of AMA with Ce(IV), Mn(III), or Fe(phen)33+ ion was studied under aerobic or anaerobic conditions. The order of reactivity of metal ions toward reaction with AMA is Fe(phen)33+ > Mn(III) > Ce(IV) under aerobic conditions whereas it is Mn(III) > Ce(IV) > Fe(phen)33+ under anaerobic conditions. Under aerobic or anaerobic conditions, the order of reactivity of RCH(CO2H)2 (R = H (MA), Me (MeMA), Et (EtMA), allyl (AMA), n‐Bu (BuMA), Ph (PhMA), and Br (BrMA)) is PhMA > MA > BrMA > AMA > MeMA > EtMA > BuMA toward reaction with Ce(IV) ion and it is MA > PhMA > BrMA > MeMA > AMA > EtMA > BuMA toward reaction with Mn(III) ion. Under aerobic conditions, the order of reactivity of RCH(CO2H)2 toward reaction with Fe(phen)33+ ion is PhMA > BrMA > (MeMA, AMA) > (BuMA, EtMA) > MA. The experiment results are rationalized.  相似文献   

8.
Silver triflate [AgOTf] assisted de‐bromination gives [Ni(dppm/dppe/(PPh3)2) (OTf)2], which on reaction with 4,4′‐bpy and gold(I) phosphines in dichloromethane medium by the self assemble technique leads to [{(L)Ni}{(4,4‐bpy)Au(PPh3)}2](OTf)4, ( 1,2,3 ) [{(L)Ni(4,4‐bpy)}4](OTf)8, ( 4,5,6 ) [L = dppm/dppe/(PPh3)2 = diphenyl phosphino‐methane, ‐ethane, bis‐triphenylphosphine, OSO2CF3 is the triflate anion]. The maximum molecular peak of the corresponding molecule is observed in the ESI mass spectrum. Ir spectra of the complexes show ‐C=C‐, ‐C=N‐, as well as phosphine stretching. The 1H NMR spectra as well as 31P (1H)NMR suggest solution stereochemistry, proton movement, and phosphorus proton interaction. Considering all the moieties, there are a lot of carbon atoms in the molecule reflected by the 13C NMR spectrum. In the 1H‐1H COSY spectrum of the present complexes and contour peaks in the 1H?13C HMQC spectrum, we assign the solution structure and stereoretentive transformation in each step.  相似文献   

9.
《Analytical letters》2012,45(9):797-803
Abstract

A coprecipitation method for concentrating amino acids from aquatic environments with a mixed hydroxide of iron(III) and copper(II) was developed. The percent coprecipitation of glycine, alanine, glutamic acid, aspartic acid, and lysine, being independent of pH in the range of 8.5 to 9.5, was more than 95%.  相似文献   

10.
The paper describes recent results of studies on the accelerating effect of sulfur‐containing protein amino acids and water activity on multistep Bi(III) ion electroreduction at mercury electrode. The catalytic effect of methionine (Mt), cystine (CY) and cysteine (CE) was analyzed based on kinetic and thermodynamic parameters, which correlated with water activity. Investigations of adsorption of those amino acids at the electrode/solution interface provided information for the analysis of the electrical double layer and its influence on the kinetics of the electrode process. The multistep Bi(III) electroreduction process is controlled by the kinetics of active complexes formation, which precedes transfer of consecutive electrons.  相似文献   

11.
Aqueous solutions of Sc(ClO4)3,ScCl3, and Sc2(SO4)3 were studied by Ramanspectroscopy over a wide concentration range. In aqueous perchlorate solutionSc(III) occurs as an hexaaqua cation. The weak, polarized Raman band assignedto the 1(a 1g) ScO6 mode of the hexaaqua-Sc (III) ion has been studied as afunction of concentration and temperature. The 1(a 1g) ScO6 mode at 442 cm–1of the hexaaqua—Sc(III) shifts only 3 cm–1 to lower frequency and broadensabout 20 cm–1 for a 60°C temperature increase. The Raman spectroscopic datasuggest that the hexaaqua-Sc (III) ion is stable in perchlorate solution within theconcentration and temperature range measured. Besides the polarized componentat 442 cm–1, two weak depolarized modes at 295 and 410 cm–1 were measuredin the Raman effect. These two modes of the ScO6 unit were assigned to 3(f 2g)and 2(e ), respectively. The infrared active mode 3(f 1u) was measured at 460cm–1. The frequency data confirm the centrosymmetry of the Sc(III) aquacomplex, contrary to earlier Raman results. The powder spectrum of crystallineSc(ClO4) 3 · 6H2O shows the above described Raman modes as well. Thesefindings are in contrast to Sc2(SO4)3 solutions, where sulfate replaces water inthe first hydration sphere and forms thermodynamically strong sulfato complexes.In ScCl3 solutions thermodynamically weak chloro complexes could be detected.Ab initio molecular orbital calculations were performed at the HF and MP2 levelsof theory using different basis sets up to 6–31 + G(d). Gas-phase structures,binding energies, and enthalpies are reported for the Sc3+(OH2)6 and Sc3+(OH2)7cluster. The Sc—O bond length for the Sc3+(OH2)6 cluster reproduces theexperimentally determined bond length of 2.18 Å (recent EXAFS data) almost exactly.The theoretical binding energy for the hexaaqua Sc(III) ion was calculated andaccounts for ca. 54–59% of the experimental hydration enthalpy of Sc(III). Thethermodynamic stability of the Sc3+(OH2)6(OH2) cluster was compared to thatof the Sc3+(OH2)7 cluster, demonstrating that hexacoordination is inherently morestable than heptacoordination in the scandium (III) system. The calculated 1ScO6frequency of the Sc+(OH2)6 cluster is ca. 12% lower than the experimentalfrequency. Adding an explicit second hydration sphere to give Sc3+ (OH2)18,denoted Sc[6 + 12], is shown to correct for the discrepancy. The frequencycalculation and the thermodynamic parameters for the Sc[6 + 12] cluster aregiven and the importance of the second hydration sphere is stressed. Calculatedfrequencies of the ScO6 subunit in the Sc[6 + 12] cluster agree very well withthe experimental values (for example, the calculated 1ScO6 frequency was foundto be 447 cm–1, in excellent agreement with the above-reported experimentalvalue). The binding enthalpy for the Sc[6 + 12]cluster predicts the single ionhydration enthalpy to about 89%.  相似文献   

12.
The kinetics of oxidation of the chromium(III)‐guanosine 5‐monophosphate complex, [CrIII(L)(H2O)4]3+(L = guanosine 5‐monophosphate) by periodate in aqueous solution to CrVI have been studied spectrophotometrically over the 25–45 °C range. The reaction is first order with respect to both [IO4?] and [CrIII], and increases with pH over the 2.38–3.68 range. Thermodynamic activation parameters have been calculated. It is proposed that electron transfer proceeds through an inner‐sphere mechanism via coordination of IO4? to chromium(III).  相似文献   

13.
Antimony(III) complexes of thioamides [thioamides=thiourea (Tu), N,N′‐dimethylthiourea (Dmtu), tetramethylthiourea (Tmtu), imidazolidine‐2‐thione (Imt) and diazinane‐2‐thione (Diaz)] with the general formulae, Sb(thione)nCl3 (n=1, 2, 2.5, 3) were prepared and characterized by elemental analysis, IR and NMR (1H, 13C) spectroscopic methods. The spectral data of the complexes are consistent with the coordination of the thiones to antimony(III). The crystal structure of one of them, {[Sb(Imt)2Cl2]2(μ2‐Imt)}Cl2 ( 1 ), was determined by X‐ray crystallography, which shows that the complex is dinuclear consisting of two [Sb(Imt)2Cl2] units bridged by an Imt molecule. In 1 , the antimony atom is bonded to two chlorine atoms, two sulfur atoms of coordinated Imt molecules and one sulfur atom of a bridging Imt molecule. The antimony environment can be considered to be distorted octahedral with one Cl? ion weakly bound to antimony.  相似文献   

14.
二(2-乙基己基)磷酸对氨基酸的萃取平衡   总被引:3,自引:0,他引:3  
二(2-乙基己基)磷酸对氨基酸的萃取平衡曹汉瑾,王德宝,刘沛妍,吴子生,严忠(东北师范大学化学系,长春,130024)关键词氨基酸,二(2-乙基己基)磷酸,萃取平衡,分配比迄今为止,有关氨基酸溶剂萃取的文献报道还不多[1~3].本实验以二(2-乙基己...  相似文献   

15.
Even though the Dewar–Chatt–Duncanson model has been successfully used by chemists since the 1950s, no experimental methodology is yet known to unambiguously estimate the constituents (donation and back‐donation) of a metal–ligand interaction. It is demonstrated here that one of these components, the metal‐to‐ligand π back‐donation, can be effectively probed by NMR measurements aimed at determining the rotational barrier of a C?N bond (ΔHr) of a nitrogen acyclic carbene ligand. A large series of gold(I) complexes have been synthesized and analyzed, and it was found that the above experimental observables show an accurate correlation with back‐donation, as defined theoretically by the appropriate charge displacement originated upon bond formation. The proposed method is potentially of wide applicability for analyzing the ligand effect in metal catalysts and guiding their design.  相似文献   

16.
Pseudo‐ephedrine derived 2‐imino‐1,3‐thiazolidine 1 reacts with tris(diethylamino)phosphane by stepwise replacement of the diethylamino group to give the mono‐, bis‐ and tris(imino)phosphanes 2 , 3 and 4 , respectively, of which 4 could be isolated in pure state. The analogous reaction with diethylamino‐diphenylphosphane affords the imino‐diphenylphosphane 5 . The iminophosphanes react with sulfur or selenium to give the corresponding phosphorus(V) compounds. In contrast, the reaction of the iminophosphanes with oxygen is very slow; anhydrous trimethylamine N‐oxide reacts in the melt with the phosphanes to give the oxides 4(O) and 5(O) . The molecular structures of 4(O) (in mixture with 4 ), 4(Se) , 5(S) and 5(Se) were determined by X‐ray analysis. In all cases the ring‐sulfur and the phosphorus atoms are in cis‐positions at the C=N bonds. The analogous solution structures were determined by 1H, 13C, 15N, 31P and 77Se NMR spectroscopy. In the case of the compounds 5 , 5(O) , 5(S) and 5(Se) the isotope‐induced chemical shifts 1δ14/15N(31P) were determined, using INEPT‐HEED experiments.  相似文献   

17.
New oxygen‐rich compounds starting from the amino acids β‐alanine and L ‐aspartic acid were synthesized and comprehensively analyzed including multinuclear NMR spectroscopy and vibrational spectroscopy. Thermal stabilities were measured and the behavior towards external stimuli like friction or impact were determined. Detonation and combustion parameters were predicted by using the EXPLO5 V6.02 code and were compared with common explosives. In addition, crystal structures were obtained for two compounds.  相似文献   

18.
Summary. The kinetics of ruthenium(III) catalysed oxidation of sulfanilic acid (p-aminobenzenesulfonic acid) by hexacyanoferrate(III) in alkaline medium at a constant ionic strength of 2.5mol·dm–3 has been studied spectrophotometrically using a rapid kinetic accessory. The reaction exhibits 2:8 stoichiometry (SNA:HCF(III)). The reaction showed first order kinetics in [hexacyanoferrate(III)] and [ruthenium(III)] and apparent less than unit order in both sulfanilic acid and alkali concentrations. The reaction rate increases with increasing ionic strength but the relative permittivity (T) of the medium has a negligible effect on the rate of the reaction. Initial addition of reaction products did not affect the rate significantly. A mechanism involving the formation of a complex between sulfanilic acid and hydroxylated species of ruthenium(III) has been proposed. The active species of HCF(III) and ruthenium(III) are understood as [Fe(CN)63–] and [Ru(H2O)5OH]2+, respectively. The main products were identified by IR, NMR, and mass spectral studies. The reaction constants involved in the different steps of mechanism are calculated. The activation parameters with respect to the slow step of the mechanism are computed and discussed and thermodynamic quantities are also calculated.  相似文献   

19.
A new complex [Pr(bib)2(NO3)3] ( 1 ) was synthesized by reaction of bidentate imidazole‐containing ligand 1‐bromo‐3,5‐bis(imidazol‐1‐ylmethyl)benzene (bib) with Pr(NO3)·6H2O and characterized by X‐ray crystallography. Complex 1 has a two‐dimensional herringbone‐like structure with the ligand bib serving as a bridging ligand using its two imidazolyl nitrogen atoms. Ligand bib adopts cis and trans two different conformations, and the Pr(III) atoms are bridged by bib in two different ways. Thermogravimetric analysis for complex 1 was carried out and the result shows that the complex is stable up to 180 °C. Variable‐temperature magnetic susceptibility of complex 1 was measured between 1.8 and 300 K and the result shows that the χMT value decreases continuously over the whole temperature range. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

20.
Salens, derived from 1,2‐ethylenediamine and salicylaldehydes, have been widely used as ligands for metal complexes which have been showing enormous potential in chemical properties of asymmetric catalysts as well as biological properties such as anticancer agents. Almost all of the salen–metal complexes with their corresponding metal (II)‐complexes show the evidences of chelation of two oxygens in salens. However, several metal (II) complexes, especially cobalt (II) complexes, could not show NMR spectra due to their paramagnetism. Recently, it has been reported that one of the cobalt (III) complexes was used for NMR spectroscopy to evaluate its stereoselectivity as a catalyst. Even though many salen ligands are known, their NMR data are not assigned completely. It was possible that modification in northern part of salen with 2‐hydroxyphenyl group afforded another oxygen chelation site in salen ligand. Here we report that synthesis and full NMR assignment of new salen ligands, which form meso 1,2‐bis(2‐hydroxyphenyl)ethylenediamine) and their cobalt (III) complexes. The assignments of 1H and 13C NMR data obtained in this experiment can help us to predict the NMR data of other salen ligands. Copyright © 2008 John Wiley & Sons, Ltd.  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