共查询到20条相似文献,搜索用时 15 毫秒
1.
The colour reactions of three oximes with ammine- and nitrosyl-pentacyanoferrate(II) ions have been studied. The optimum reaction conditions, spectral characteristics and molar compositions have been found and some instability constants determined. The reactions can be used for the detection and determination of small amounts of the oximes examined. 相似文献
2.
The pertechnetate ion oxidizes ascorbic acid in strong acid medium to form red species. A reaction mechanism has been developed which correctly predicts all the experimental facts. The results obtained support the postulate according to which the red species corresponds to a complex formed between Tc(V) and dehydroascorbic acid. The rate constants and Arrhenius parameters have been investigated. 相似文献
3.
Hasan Marai Ewa Kita Joanna Wojciechowska Paula Wróbel 《Transition Metal Chemistry》2012,37(2):215-223
The following chromium(III) complexes with serine (Ser) and aspartic acid (Asp) were obtained and characterized in solution:
[Cr(ox)2(Aa)]2− (where Aa = Ser or Asp), [Cr(AspH−1)2]− and [Cr(ox)(Ser)2]−. In acidic solutions, [Cr(ox)2(Aa)]2− undergoes acid-catalysed aquation to cis-[Cr(ox)2(H2O)2]− and the appropriate amino acid. [Cr(ox)(Ser)2]− undergoes consecutive acid-catalysed Ser liberation to give [Cr(ox)(H2O)4]+, and the [Cr(Asp)2]− ion is converted into [Cr(Asp)(H2O)4]2+. Kinetics of these reactions were studied under isolation conditions. The determined rate expressions for all the reactions
are of the form: k
obs = a + b[H+]. Reaction mechanisms are proposed, and the meaning of the determined parameters has been established. Evidence for the formation
of an intermediate with O-monodentate amino acid is given. The effect of the R-substituent at the α-carbon atom of the amino
acid on the complex reactivity is discussed. 相似文献
4.
S. Shyamroy B. Garnaik S. Sivaram 《Journal of polymer science. Part A, Polymer chemistry》2005,43(10):2164-2177
The synthesis of low‐molecular‐weight (weight‐average molecular weight < 45,000 g/mol) lactic acid polymers through the dehydropolycondensation of L ‐lactic acid was investigated. Polymerizations were carried out in solution with solvents (xylene, mesitylene, and decalin), without a solvent using different Lewis acid catalysts (tetraphenyl tin and tetra‐n‐butyldichlorodistannoxane), and at three different polymerization temperatures (143, 165, and 190 °C). The products were characterized with differential scanning calorimetry, size exclusion chromatography, vapor pressure osmometry, 13C NMR, and matrix‐assisted laser desorption/ionization time‐of‐flight (MALDI‐TOF). The resulting polymers contained less than 1 mol % lactide, as shown by NMR. The number‐average molecular weights were calculated from the ratio of the area peaks of ester carbonyl and carboxylic acid end groups via 13C NMR. The stereosequences were analyzed by 13C NMR spectroscopy on the basis of triad effects. Tetraphenyl tin was an effective transesterification catalyst, and the randomization of the stereosequence at 190 °C was observed. In contrast, the distannoxane catalyst caused comparatively less transesterification reaction, and the randomization of the stereosequences was slow even at 190 °C. The L ‐lactic acid and D ‐lactic acid isomers were added to the polymer chain in a small, blocky fashion. The MALDI‐TOF spectra of poly(L ‐lactic acid) (PLA) chains doped with Na+ and K+ cations showed that the PLA chains had the expected end groups. The MALDI‐TOF analysis also enabled the simultaneous detection of the cyclic oligomers of PLA present in these samples, and this led to the full structural characterization of the molecular species in PLA. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 2164–2177, 2005 相似文献
5.
The reactions of platinum(II) complexes, [PtCl2(dach)] (dach = (1R,2R)‐1,2‐diaminocyclohexane) and [PtCl2(en)] (en = ethylenediamine) with biologically relevant ligands such as 5′‐GMP (guanosine‐5′‐monophosphate) and l ‐His (l ‐histidine) were studied by UV–vis spectrophotometry, 1H NMR spectroscopy, and high‐performance liquid chromatography (HPLC). Spectrophotometrically, these reactions were investigated under pseudo‐first‐order conditions at 310 K in 25 mM Hepes buffer (pH 7.2) and 10 mM NaCl to prevent the hydrolysis of the complexes. The [PtCl2(en)] complex reacts faster than [PtCl2(dach)] in the reaction with studied nucleophiles. This confirms the fact that the reactivity of studied Pt(II) complexes depends on the structure of the inert bidentate ligand. Also, the substitution reactions with l ‐His are always faster than the reactions with nucleotide 5′‐GMP. The reactions of [PtCl2(dach)] and [PtCl2(en)] complexes with l ‐histidine are studied by 1H NMR spectroscopy. The obtained rate constants are in agreement with those obtained by UV–vis. The same reactions were studied by HPLC comparing the obtained chromatograms during the reaction. The changes in intensity of signals of the free and coordinated ligand show that after a few days there is only one dominant product in the system. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 43: 99–106, 2011 相似文献
6.
Sai Chodavarapu Seshatalpa Hela Pamidipati Gayatri Sridhar Yeleswarapu Anipindi Nageswara Rao 《Transition Metal Chemistry》1998,23(3):249-252
The reaction of the Cr(xx)2(H2O)2
− (xx = oxalate, malonate and methylmalonate) complexes with dissolved CO2 was studied by stopped-flow spectrophotometry in the 7 < pH < 9 range and between 20 to 30°C at an ionic strength of 0.5
mol dm−3 (NaCl). Under the experimental conditions the aqua complex ion consists of a pH-dependent mixture of Cr(xx)2(H2O)2
−, Cr(xx)2(OH) (H2O)2− and Cr(xx)2(OH)2
3−. The monohydroxo and dihydroxo species undergo CO2 uptake and subsequent intramolecular carbonate ligand chelation independently, at rates which are readily distinguishable
and are governed by the uptake rate constants k
1 and k
2 and chelation rate constants k
3 and k
4, respectively. Only the k
1 values for oxalato, malonato and methylmalonato complexes could be calculated; k
1 = 1084 and 1333 and 1650 mol−1 dm3 s−1, respectively. The results obtained were compared with those obtained from other systems that have either cobalt(III), iridium(III)
or rhodium(III) as central atoms.
This revised version was published online in June 2006 with corrections to the Cover Date. 相似文献
7.
Chang‐Ming Dong Kun‐Yuan Qiu Zhong‐Wei Gu Xin‐De Feng 《Journal of polymer science. Part A, Polymer chemistry》2002,40(3):409-415
Two types of three‐arm and four‐arm, star‐shaped poly(D,L ‐lactic acid‐alt‐glycolic acid)‐b‐poly(L ‐lactic acid) (D,L ‐PLGA50‐b‐PLLA) were successfully synthesized via the sequential ring‐opening polymerization of D,L ‐3‐methylglycolide (MG) and L ‐lactide (L ‐LA) with a multifunctional initiator, such as trimethylolpropane and pentaerythritol, and stannous octoate (SnOct2) as a catalyst. Star‐shaped, hydroxy‐terminated poly(D,L ‐lactic acid‐alt‐glycolic acid) (D,L ‐PLGA50) obtained from the polymerization of MG was used as a macroinitiator to initiate the block polymerization of L ‐LA with the SnOct2 catalyst in bulk at 130 °C. For the polymerization of L ‐LA with the three‐arm, star‐shaped D,L ‐PLGA50 macroinitiator (number‐average molecular weight = 6800) and the SnOct2 catalyst, the molecular weight of the resulting D,L ‐PLGA50‐b‐PLLA polymer linearly increased from 12,600 to 27,400 with the increasing molar ratio (1:1 to 3:1) of L ‐LA to MG, and the molecular weight distribution was rather narrow (weight‐average molecular weight/number‐average molecular weight = 1.09–1.15). The 1H NMR spectrum of the D,L ‐PLGA50‐b‐PLLA block copolymer showed that the molecular weight and unit composition of the block copolymer were controlled by the molar ratio of L ‐LA to the macroinitiator. The 13C NMR spectrum of the block copolymer clearly showed its diblock structures, that is, D,L ‐PLGA50 as the first block and poly(L ‐lactic acid) as the second block. © 2001 John Wiley & Sons, Inc. J Polym Sci Part A: Polym Chem 40: 409–415, 2002 相似文献
8.
9.
《Polyhedron》1986,5(3):663-665
The colour reactions of three bis-pyridinium aldoximes with ammine- and nitrosyl-pentacyanoferrate(II) ions were studied. The optimum reaction conditions, the compositions and the stability constants of the produced complexes were determined. The compelxes were isolated and characterized by elemental analysis, visible, UV and IR spectral data. The examined compounds, although dioximes, are found to form 1 : 1 complexes. 相似文献
10.
Noemí D. Lis de Katz Néstor E. Katz 《Monatshefte für Chemie / Chemical Monthly》1982,113(6-7):745-750
The preparation and spectral and kinetic properties of new complexes of pentacyanoferrate(II) coordinated to methyl and ethyl 4-pyridinecarboxylates are described. Ester substitution by solvent water seems to compete effectively with the alkaline or acid ester hydrolysis due to a remarkable metal inhibition of the latter reaction.Presented in part at the XVas. Sesiones Químicas Argentinas, Horco Molle, Tucumán, Argentina, september 1980. 相似文献
11.
《Journal of Inorganic and Nuclear Chemistry》1976,28(3):431-434
The intervalence spectra of several iron(III) and copper(II) pentacyanoferrate(II) complexes have been studied in comparison to those of the hexacyanoferrate(II) analogs. The energies of the intervalence transitions have shown to be strongly dependent on the properties of the Fe(CN)5Ln− anion. The observed trend on the energies of the intervalence bands in the sequence NH3 < py < pyrazinecarboxylate < dimethyl sulfoxide ∼ tetramethylene sulfoxide < CO, for the ligand L, were interpreted in terms of backbonding stabilization of the low spin, iron(II) donor ion. 相似文献
12.
Olayinka A. Oyetunji Banyaladzi D. Paphane Clifford A. L. Becker 《Transition Metal Chemistry》2006,31(7):951-957
The interaction of pyridine with four tetrakis(arylisocyanide)cobalt(II) complexes, [Co(CNR)4(ClO4)2] R = 2,6-Me2C6H3 (A), 2,4,6-Me3C6H2 (B), 2,6-Et2C6H3 (C) and 2,6-iPr2C6H3 (D), have been studied in 2,2,2-trifluoroethanol medium. The kinetics of the reactions were investigated over the 293–318 K
temperature range. The reaction profile exhibited two distinct processes, proposed to be an initial fast substitution followed
by a slow reduction, for each of the reactions. The pseudo first-order rate constants for both processes increased with increasing
concentration of pyridine with the reduction processes exhibiting saturation kinetics at high pyridine concentrations. Steric
hindrance plays a significant role in the rates of the reactions, as the rates decrease in the order k(A) > k(B) > k(C) > k(D). The activation enthalpies, ΔH‡, increase from A to D while the activation entropies, ΔS‡, are relatively similar for the four reactions, indicating similar transition states and hence similar mechanisms. Complex
B was first synthesized and characterized in this study. 相似文献
13.
The proposed method for ascorbic acid: AA (Vitamin C) determination is based on the oxidation of AA to dehydroascorbic acid with the CUPRAC reagent of total antioxidant capacity assay, i.e., Cu(II)-neocuproine (Nc), in ammonium acetate-containing medium at pH 7, where the absorbance of the formed bis(Nc)-copper(I) chelate is measured at 450 nm. The flavonoids (essentially flavones and flavonols) normally interfering with the CUPRAC procedure were separated with preliminary extraction as their La(III) chelates into ethylacetate (EtAc). The Cu(I)-Nc chelate responsible for color development was formed immediately with AA oxidation. Beer's law was obeyed between 8.0 × 10−6 and 8.0 × 10−5 M concentration range, with the equation of the linear calibration curve: A450 nm = 1.60 × 104C (mol dm−3) − 0.0596. The relative standard deviation (R.S.D.) in the analysis of N = 45 synthetic mixtures containing 1.25 × 10−2 mM AA with flavonoids was 5.3%. The Cu(II)-Nc reagent is a lower redox-potential and therefore more selective oxidant than the Fe(III)-1,10-phenanthroline reagent conventionally used for the same assay. This feature makes the proposed method superior for real samples such as fruit juices containing weak reductants such as citrate, oxalate and tartarate that may otherwise produce positive errors in the Fe(III)-phen method when equilibrium is achieved. The developed method was applied to some commercial fruit juices and pharmaceutical preparations containing Vitamin C + bioflavonoids. The findings of the developed method for fruit juices and pharmaceuticals were statistically alike with those of HPLC. The proposed spectrophotometric method was practical, low-cost, rapid, and could reliably assay AA in the presence of flavonoids without enzymatic procedures open to interferences by enzyme inhibitors. 相似文献
14.
A new, simple and rapid method of determination of ascorbic acid in amounts of 45-360 mug is described. The ascorbic acid is determined spectrophotometrically at 420 nm from the decrease in absorbance it causes in 1 x 10(-3)M hexacyanoferrate(III) in McIlvaine buffer at pH 5.2. The proposed method is suitable for the determination of ascorbic acid in pharmaceutical preparations and probably natural products. 相似文献
15.
Keiichi Taki Ichiro Arita Mitsuru Satoh Jiro Komiyama 《Journal of Polymer Science.Polymer Physics》1999,37(10):1035-1041
Permeability coefficients of D ‐ and L ‐tryptophan (D ‐, L ‐Trp) were estimated for poly(L ‐glutamic acid) (PLG) membranes immersed in aqueous ethanol. D ‐tryptophan was selectively transported (the maximum permeability ratio was 2.6) depending on the amount and the species of crosslinking agent, and on the composition of immersing solvent. It is suggested that hydrogen bonding between uncharged permeates and carboxyl and/or amide groups of PLG is an essential factor for the selective transport. © 1999 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 37: 1035–1041, 1999 相似文献
16.
Hasan Marai Ewa Kita Emilia Kiersikowska Sylwia Kuchta Anna Bajek Tomasz Drewa 《Transition Metal Chemistry》2012,37(4):337-344
Three chromium(III) complexes with asparagine (Asn) and histidine (His) of the [Cr(ox)2(Aa)]2− type, where Aa = N,O–Asn, N,O–His or N,N′–His, were obtained and characterized in solution. The complexes with N,O–Aa undergo acid-catalysed aquation to give a free amino acid and cis-[Cr(ox)2(H2O)2]−, whereas the complex with N,N′–His undergoes parallel reaction paths: (1) isomerization to the N,O–His complex and (2) liberation of an oxalate ligand. Kinetics of the N,O–Aa complexes in HClO4 media were studied spectrophotometrically under pseudo-first-order conditions. The absorbance changes were attributed to
the chelate ring opening at the Cr–N bond. The linear dependence of rate constants on [H+] was established, and a mechanism for the chelate ring cleavage was postulated. The existence of a metastable intermediate
with O-monodentate Aa ligand was proved experimentally. Effect of [Cr(ox)2(Aa)]2− on 3T3 fibroblasts proliferation was studied. The tests revealed low cytotoxicity of the complexes. Complexes with Ala, His
and Cys are good candidates for biochromium sources. 相似文献
17.
Joanna Wiśniewska 《Transition Metal Chemistry》2007,32(6):811-815
The kinetics of the electron-transfer reactions between promazine (ptz) and [Co(en)2(H2O)2]3+ in CF3SO3H solution ([CoIII] = (2–6) × 10−3
m, [ptz] = 2.5 × 10−4
m, [H+] = 0.02 − 0.05 m, I = 0.1 m (H+, K+, CF3SO
3
−
), T = 288–308 K) and [Co(edta)]− in aqueous HCl ([CoIII] = (1 − 4) × 10−3
m, [ptz] = 1 × 10−4
m, [H+] = 0.1 − 0.5 m, I = 1.0 m (H+, Na+, Cl−), T = 313 − 333 K) were studied under the condition of excess CoIII using u.v.–vis. spectroscopy. The reactions produce a CoII species and a stable cationic radical. A linear dependence of the pseudo-first-order rate constant (k
obs) on [CoIII] with a non-zero intercept was established for both redox processes. The rate of reaction with the [Co(en)2(H2O)2]3+ ion was found to be independent of [H+]. In the case of the [Co(edta)]− ion, the k
obs dependence on [H+] was linear and the increasing [H+] accelerates the rate of the outer-sphere electron-transfer reaction. The activation parameters were calculated as follows:
ΔH
≠ = 105 ± 4 kJ mol−1, ΔS
≠ = 93 ± 11 J K−1mol−1 for [Co(en)2(H2O)2]3+; ΔH
≠ = 67 ± 9 kJ mol−1, ΔS
≠ = − 54 ± 28 J K−1mol−1 for [Co(edta)]−. 相似文献
18.
Matthew R. Wood Thomas A. Brettell Roger A. Lalancette 《Acta Crystallographica. Section C, Structural Chemistry》2007,63(2):m33-m35
The title salt, methyl (1R,2R,3S,5S,8S)‐3‐benzoyloxy‐8‐methyl‐8‐azabicyclo[3.2.1]octane‐2‐carboxylate tetrachloroaurate(III), (C17H22NO4)[AuCl4], has its protonated N atom intramolecularly hydrogen bonded to the O atom of the methoxycarbonyl group [N⋯O = 2.755 (6) Å and N—H⋯O = 136°]. Two close intermolecular C—H⋯O contacts exist, as well as five C—H⋯Cl close contacts. The [AuCl4]− anion was found to be distorted square planar. 相似文献
19.
The kinetic study and mechanism of the permanganic oxidation of L‐glutamine in sulfuric acid has been carried out both in the absence and presence of silver (I) using a spectrophotometric technique. Activation parameters have been evaluated using the Arrhenius and Eyring plots. Mechanisms consistent with the observed kinetic data have been proposed and discussed. The overall rate expression for the oxidation may be written as In the presence of silver (I) the rate law is The reaction appears to involve an acid catalyzed and data showed role of water molecules in the rate‐determining step is proton transfer which satisfies Bunnett's theory. A mechanism satisfying the various kinetic parameters has been proposed. ©1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 95–102, 1999 相似文献
20.
Martina Osti Michael F. Cunningham Ralph Whitney Barkev Keoshkerian 《Journal of polymer science. Part A, Polymer chemistry》2007,45(1):69-80
Styrene miniemulsions that were stabilized by common anionic surfactants (sodium dodecyl benzene sulfonate, sodium dodecyl sulfate, and disulfonated alkyl diphenyl oxide sodium salt) polymerized at 25 °C in the presence of L ‐ascorbic acid without the addition of a free‐radical initiator. The polymerizations exhibited high rates and molecular weights, with conversions greater than 70% achieved in less than 1 h and weight‐average molecular weights greater than 1 × 106 g/mol. Polymers did not form in the absence of L ‐ascorbic acid. Although the final conversion was only slightly independent on the L ‐ascorbic acid concentration, it was dependent on the surfactant concentration. The rate and final conversion were also strongly dependent on the surfactant type. Moreover, it was possible to initiate polymerizations with a monomer‐soluble derivative of L ‐ascorbic acid (L ‐ascorbic acid 6‐palmitate), although the rates were dramatically reduced compared with those when water‐soluble L ‐ascorbic acid was used. High yields and high‐molecular‐weight polymers were also produced with butyl acrylate and methyl methacrylate with L ‐ascorbic acid in the presence of sodium dodecyl benzene sulfonate. The initiation was attributed to an interaction between the surfactant and L ‐ascorbic acid, which formed a redox initiation system that generated radicals capable of adding monomer. These results are of particular significance for redox‐initiated emulsion/miniemulsion polymerizations with L ‐ascorbic acid as the reducing agent and with sulfate or sulfonate surfactants. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 45: 69–80, 2007 相似文献