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 31 毫秒
1.
The nonenzymatic reaction of ethanol-derived CH3CHO with tissue constituents continues to be of interest as a potential mechanism underlying the toxicity of alcohol. The current study has focused on the spontaneous condensation of CH3CHO with H4folate under physiological conditions (38 °C, pH 7.0, I = 0.25 M). Computer analysis of uv spectral changes with increasing CH3CHO concentrations demonstrated the presence of at least two different adducts. The observed equilibrium constant (Kobs) for the formation of the first adduct is 91 ± 2 m?1 (121 ± 2 m?1 at 25 °C), a value which is unaffected by variations in ionic strength (0.06–1.0 m) or by free [Mg2+] up to 5 mm. The NMR spectrum is compatible with the structure: 5,10-CH3CH-H4folate analogous to the naturally occurring 5,10-CH2-H4folate. The formation of the latter compound from HCHO and H4folate, however, is much more favorable under the same conditions [Kobs = 3.0 ± 0.2 × 104 M?1 (38 °C), 3.6 ± 0.1 × 104 M?1 (25 °C)]. At the levels of CH3CHO which accumulate during ethanol metabolism in vivo only a small fraction of the H4folate will exist as the CH3CHO derivative, yet it may ultimately be the ratio of free CH3CHO to free HCHO in tissue which determines the physiological importance of the CH3CHO adduct. Other adduct(s) of CH3CHO with H4folate are observed at very high levels of CH3CHO but are unlikely to be of physiological significance.  相似文献   

2.
《Free radical research》2013,47(3-6):381-388
The abilities of chemically generated hydroxyl radical (OH), superoxide anion (O?) and hydrogen peroxide (H2O2) to degrade rat myocardial membrane phospholipids previously lableed with [1 -14C]arachidonic acid were studied. HO and H2O2 but not O2??, caused the degradation of phospha-tidylcholine (PC), phosphatidylethanolamine (PE), and phosphatidylinositol (PI). With OH' and H2O2, the loss of radiolable in PC was accompanied by an increase in the radiolabel of lysophosphatidylcholine (LPC), but not in that of free fatty acid (FFA). These results suggest the hydrolysis of l-oxygen ester bond of PC by HO' and that H2O2 and that HO' and H2O2, but not O?, are detrimental to the structure and function of membrane phospholipids. However, since μM amounts of HO' and mM amounts of H2O2 were necessary to affect the membrane phospholipids, it is likely that in the reprefused myocardium only HO', but not H2O2, may directly cause the breakdown of membrane phospholipids.  相似文献   

3.
Given the increase in the incidence of insulin resistance, obesity, and type 2 diabetes in children and adolescents, it would be of paramount importance to assess quantitative indices of insulin secretion and action during a physiological perturbation, such as a meal or an oral glucose‐tolerance test (OGTT). A minimal model method is proposed to measure quantitative indices of insulin secretion and action in adolescents from an oral test. A 7 h, 21‐sample OGTT was performed in 11 adolescents. The C‐peptide minimal model was identified on C‐peptide and glucose data to quantify indices of β‐cell function: static φs and dynamic φd responsivity to glucose from which total responsivity φ was also measured. The glucose minimal model was identified on glucose and insulin data to estimate insulin sensitivity, SI, which was compared to a reference measure, SIref, provided by a tracer method. Disposition indices, which adjust insulin secretion for insulin action, were then calculated. Indices of β‐cell function were φs = 51.35 ± 8.89 × 10?9min?1, φd = 1,392 ± 258 × 10?9, and φ = 82.09 ± 17.70 × 10?9min?1. Insulin sensitivity was SI = 14.19 ± 2.73 × 10?4, not significantly different from SIref = 14.96 ± 3.04 × 10?4 dl/kg·min per µU/ml, and well correlated: r = 0.98, P < 0.0001, thus indicating that SI can be accurately measured from an oral test. Disposition indices were DIs = 1,040 ± 201 × 10?14 dl/kg/min2 per pmol/l, DId = 33,178 ± 10,720 × 10?14 dl/kg/min per pmol/l, DI = 1,844 ± 522 × 10?14 dl/kg/min2 per pmol/l. Virtually the same minimal model assessment was obtained with a reduced 3 h, 9‐sample protocol. OGTT interpreted with C‐peptide and glucose minimal model has the potential to provide novel insight regarding the regulation of glucose metabolism in adolescents, and to evaluate the effect of obesity and interventions such as diet and exercise.  相似文献   

4.
Biogenic emissions of halomethanes (CH3CI, CH3Br and CH3I) and methanethiol (CH3SH) are of major significance to atmospheric chemistry, but there is little information on such emissions from higher plants. We present evidence that plants can produce all these gases through an identical methyltransferase reaction. A survey of 118 herbaceous species, based on CH3I production by leaf discs supplied with Kl, detected the presence of in vivo halide methyltransferase activity in 87 species. The activities ranged over nearly 4 orders of magnitude. Plants generally considered salt tolerant had relatively low activities, and salinization of three such species did not increase the activity. The highest activities were found in the family Brassicaceae. Leaf extracts of Brassica oleracea catalysed the S-adenosyl-L-methioninc-dependenl niethylalion of the halides I?, Br? and CI? to the respective halomethanes. In addition, the extract similarly methylated HS? (bisulphide) to CH3SH. These two types of enzyme activity (halide and bisulphide methyltransferase) were also present in all of the 20 species comprising a subsample that represented the range of CH3I emissions observed in the initial survey of in vivo CH3I production ability, and in a marine red alga Endocladia muricata. Moreover, the two activities occurred in approximately the same ratio in all the higher plants tested. These findings highlight the potential of higher plants to contribute to the atmosphericbudget of halomethanes and melhanethiol. The halide and bisulphide methyltransferase activities may also provide a mechanism for the elimination of halide and HS? ions, both of which are known to be phytotoxic.  相似文献   

5.
Addition of methyl-coenzyme M (CH3SCH2CH2SO3?) to undialized, anaerobic, cell-extracts of Methanobacterium thermoautotrophicum under an atmosphere of H2 and CO2 (80:20 v/v) stimulates 30-fold the rate of CO2 reduction to methane. For each mol of CH3SCH2CH2SO3? added 12 mol of methane is produced. This stimulation phenomenon requires magnesium ion, ATP, H2, and CH3SCH2CH2SO3?. Neither the reduced form of the cofactor, HSCH2CH2SO3?, nor the oxidized, disulfide form will replace the methylated coenzyme.  相似文献   

6.
ABSTRACT

Monomethylmercury (CH3Hg +) is both the most ecologically significant and the least well characterized species of mercury in environmental settings. Our understanding of the environmental speciation behavior of this compound is limited both as the result of lesser available laboratory data (when compared to inorganic mercury) as well as the uncertainties associated with our understanding of the properties of environmental ligands. A careful examination and synthesis of data reported in the technical literature led to the following findings: (1) a 25°C, zero ionic strength bicarbonate ion complexation constant estimate is remarkably close to an earlier reported value at 0.4 M: CH3Hg+ + HCo3-?CH3HgHCO3,log10K = 2.6 (±0.22, 1 SD), (2) three 25°C zero ionic strength reaction constants reported by DeRobertis et al.(1998) were confirmed to within ~±0.1 log10K units: CH3Hg ++ OH-?CH3HgOH, log10K = 9.47; 2CH3Hg + + H2O?(CH3Hg)2OH + + H+, log10K =?2.15; CH3Hg ++ Cl-?CH3HgCl, log10K = 5.45, (3) “best estimate” literature complexation constants corrected to zero ionic strength include: CH3Hg + + F-?CH3HgF, log10K = 1.75 (20°C corr. Schwart-zenbach and Schellenberg, 1965); CH3Hg + + Br-?CH3HgBr, log10K = 6.87 (20°C corr. Schwartzenbach and Schellenberg, 1965); CH3Hg + +1-?CH3HgI, log10K = 8.85 (20°C corr. Schwartzenbach and Schellenberg, 1965); and CH3Hg ++ SO42-?CH3HgSO4-,log10K = 2.64 (25°C, DeRobertis et al., 1998), (4) literature reported values for simulating monomethylmercury complexation with the carbonate ion may be too low: CH3Hg ++ CO32-?CH3HgCO3-, log10K = 6.1 (Rabenstein et al., 1976; Erni, 1981), and (5) ‘‘best estimate’’ constants for simulating methyl mercury complexation with reduced environmental sulfur species include: CH3Hg + + S2-?CH3HgS -, log10K = 21.1; CH3Hg ++ SH -? CH3HgSH, log10K = 14.5 (H + + SH-?CH2S, log10K = 6.88; Dyrssen and Wedborg, 1991); CH3Hg + + RS-?CH3HgSR, log10K = 16.5 (H + + RS-?RSH, log10K = 9.96; Qian et al., 2002); and CH3Hg ++ CH3HgS1 -?(CH3Hg)2S, log10K = 16.32 (Schwartzenbach and Schellenberg, 1965; Rabenstein et al., 1978; and Erni, 1981).  相似文献   

7.
Using density functional theory calculations, we investigated properties of a functionalized BC2N nanotube with NH3 and five other NH2-X molecules in which one of the hydrogen atoms of NH3 is substituted by X = ?CH3, ?CH2CH3, ?COOH, ?CH2COOH and ?CH2CN functional groups. It was found that NH3 can be preferentially adsorbed on top of the boron atom, with adsorption energy of ?12.0 kcal mol?1. The trend of adsorption-energy change can be correlated with the trend of relative electron-withdrawing or -donating capability of the functional groups. The adsorption energies are calculated to be in the range of ?1.8 to ?14.2 kcal mol?1, and their relative magnitude order is found as follows: H2N(CH2CH3) > H2N(CH3) > NH3 > H2N(CH2COOH) > H2N(CH2CN) > H2N(COOH). Overall, the functionalization of BC2N nanotube with the amino groups results in little change in its electronic properties. The preservation of electronic properties of BC2N coupled with the enhancement of solubility renders their chemical modification with either NH3 or amino functional groups to be a way for the purification of BC2N nanotubes.  相似文献   

8.
Lactoperoxidase, in the presence of H2O2, I?, and rat liver microsomes, will peroxidize membrane lipids, as evidence by malondialdehyde formation. Fe3+ assists in the formation of malondialdehyde. Fe3+ can be added at the end of the reaction period as well as at the beginning with equal effectiveness, suggesting that it only acts to assist in the conversion of lipid peroxides, previously formed by lactoperoxidase, to malondialdehyde. The addition of EDTA to the microsomal reaction mixture results in a 40% decrease in malondialdehyde formation. The antioxidant butylated hydroxytoluene will completely block the formation of malondialdehyde. Malondialdehyde formation is not dependent upon the production of superoxide, singlet oxygen, or hydroxyl radicals. Peroxidation of membrane lipids by this system is equally effective in both intact microsomes and in liposomes, indicating that iodination of microsomal protein is not required for lipid peroxidation to occur.  相似文献   

9.
The kinetic model of toluene decomposition in nonequilibrium low-temperature plasma generated by a pulse-periodic discharge operating in a mixture of nitrogen and oxygen is developed. The results of numerical simulation of plasma-chemical conversion of toluene are presented; the main processes responsible for C6H5CH3 decomposition are identified; the contribution of each process to total removal of toluene is determined; and the intermediate and final products of C6H5CH3 decomposition are identified. It was shown that toluene in pure nitrogen is mostly decomposed in its reactions with metastable N2(A3?? u + ) and N2(a??1?? u ? ) molecules. In the presence of oxygen, in the N2 : O2 gas mixture, the largest contribution to C6H5CH3 removal is made by the hydroxyl radical OH which is generated in this mixture exclusively due to plasma-chemical reactions between toluene and oxygen decomposition products. Numerical simulation showed the existence of an optimum oxygen concentration in the mixture, at which toluene removal is maximum at a fixed energy deposition.  相似文献   

10.
《Inorganica chimica acta》1988,148(2):255-260
Arytellurol complexes [PtCl(TeAr)(PPh3)2] (I) and [Pt(TeAr)2(PPh3)2] (II) are readily obtained from cis-[PtCl2(PPh)3)2] and NaTeAr (Ar = C6H5, 4-CH3OC6H4 and 4-CH3CH2OC6H4) in ethanolbenzene at room temperature. 31P NMR spectra of (I) and (II) indicate their trans configuration in solution. Metathetical reactions between I (Ar = 4-CH3OC6H4) and NaX (X = I, Br, SCN) occur in methanol to give [Pt(X)(TeC6H4OCH3-4)(PPh3)2]. 1H NMR shows that equimolar proportions of NaTeC6H5, NaTeC6H4OCH2CH3-4 and cis-[PtCl2(PPh3)2] give a mixture of three complexes: II, Ar = C6H5; II, Ar = 4-CH3CH2OC6H4; and [Pt(TeC6H5)(TeC6H4OCH2CH3-4)(PPh3)2]. Polymeric complexes [PtCl(TeAr)]n (III) and [Pt(TeAr)2]n (IV) result from reaction between K2[PtCl4] and NaTeAr in aqueaous ethanol. They react with excess of PPh3 in CDCl3 to yield monomeric complexes I and II respectively which were characterized in situ by 1H and 31P NMR of the reaction mixtures. IR spectra indicate the presence of bridging chloride ligands in III. An alternating chloride and tellurol bridged chain structure for III and a tellurol bridged for IV have been proposed. Reaction between equimolar amounts of III and PPh3 in dichloromethane yielded a tellurol bridged dimeric complex [PtCl(μ-TeAr)(PPh3)]2 (V) with terminal chloride ligand as suggested by IR study. Ethanolic solutions of diarylditellurides also react readily with an aqueous solution of K2[PtCl4] at 10 °C to give complexes for which the structure trans-[PtCl2(ArTeTeAr)2] (VI) is suggested from their elemental analyses, IR, Raman (in one case only), 1H, 125Te (in one case only), and 195Pt NMR spectra and reactions with triphenylphosphine which liberated free ditellurides. At 40 °C or above the same ditellurides form polymeric complexes III with K2[PtCl4] in aquaeous ethanol.  相似文献   

11.
The reaction of the dimeric zinc(II) chelates of the type I (R1 = R2 = CH3, R1 = H, R2 = Ph) with pyridine, 2-methylpyridine, 3-methylpyridine and 4-methylpyridine afforded the monomeric monobase adducts. The isolated adducts were characterized by their electronic and 1H NMR spectra, and a five coordinate square pyramidal structure was tentatively assigned for these adducts.The adduct formation reaction was followed spectrophotometrically and the reaction kinetics were studied using a stopped flow technique. From the available kinetic data, as well as the measured activated parameters (ΔH#, ΔS#), a mechanism for the adduct formation reaction is proposed.  相似文献   

12.
The 1,3-oxazine complexes cis- and trans-[PtCl2{ C(R)OCH2CH2C}H22] (cis: R=CH3 (1a), CH2CH3 (2a), (CH3)3C (3a), C6H5 (4a); trans:R =CH3 (1b), C6H5 (4b)) were obtained in 51-71% yield by reaction in THF at 0 °C of the corresponding nitrile complexes cis- and trans-[PtCl2(NCR)2] with 2 equiv. of OCH2CH2CH2Cl, generated by deprotonation of 3-chloro-1-propanol with n-BuLi. The cationic nitrile complexes trans-[Pt(CF3)(NCR)(PPh3)2]BF4 (R=CH3, C6H5) react with 1 equiv, of OCH2CH2CH2Cl to give a mixture of products, including the corresponding oxazine derivatives trans-[Pt(CF3){ CH2}(PPh3)2]BF4 (5 and 6), the chloro complex trans- [Pt(CF3)Cl(PPh3)2] and free oxazine H2. For short reaction times (c. 5–15 min) the oxazine complexes 5 and 6 could be isolated in modest yield (37–49%) from the reaction mixtures and they could be separated from the corresponding chloro complex (yield 40%) by taking advantage of the higher solubility of the latter derivative in benzene. For longer reaction times (> 2 h), trans-[Pt(CF3)Cl(PPh3)2] was the only isolated product. Complex 6 was crystallographically characterized and it was found to contain also crystals of trans- [PtCl{ H2}(PPh3)2]BF4, which prevented a more detailed analysis of the bond lengths and angles within the metal coordination sphere. The 1,3-oxazine ring, which shows an overall planar arrangement, is characterized by high thermal values of the carbon atoms of the methylene groups indicative of disordering in this part of the molecule in agreement with fast dynamic ring processes suggested on the basis of 1H NMR spectra. It crystallizes in the trigonal space group P , with a=22.590(4), b=15.970(3) Å, γ=120°, V=7058(1) Å3 and Z=6. The structure was refined to R=0.059 for 3903 unique observed (I3σ(I)) reflections. A mechanism is proposed for the conversion of nitrile ligands to oxazines in Pt(II) complexes.  相似文献   

13.
A novel Cu(II) complex chemosensor for hydrogen sulfide with azo as the colorimetric group has been synthesized. The complex and ligand crystals were obtained and the molecular structures were characterized by X‐ray diffraction and Electrospray ionization High resolution mass spectrometer (ESI‐HRMS). The photophysical and recognition properties were examined. The complex can recognize S2?, with an obvious color change from yellow to red based on a copper ion complex displacement mechanism. By contrast, no obvious changes were observed in the presence of other anions (AcO?, H2PO4?, F?, Cl?, Br? and I?). We present a simple, easily prepared, yet efficient, inorganic reaction‐based sensor for the detection of S2?. The complex should have many chemical and analytical applications in the sensing of hydrogen sulfide.  相似文献   

14.
The β-ligands     
The 5-hydroxybenzimidazolylcobamide (B12-HBI) derivatives from Methanosarcina barkeri were isolated. SO32?, CN?, H2O, NH3 and CH3? were identified as β-ligands. Two B12-HBI compounds with unidentified β-ligands were found, of which one constituted a major part of the corrinoid content. 5′-Deoxyadenosyl was found as a β-ligand of a corrinoid without α-ligand. Biosynthesis of CH3B12HBI was observed in cell-free extracts and depended on methanol and ATP.  相似文献   

15.
The lactate dehydrogenase-catalyzed reduction of pyruvate by NADH was studied using a spectroscopic method. The inhibitory effect exhibited by high concentrations of pyruvate was investigated in phosphate and 2,2-diethylmalonate buffers. Kinetic studies were carried out in which the rate of the enzyme-catalyzed reaction was monitored at various stages of pyruvate hydration, H2O + CH3COCO2? ? CH3C(OH)22C02?. Buffered solutions of different initial relative amounts of ketopyruvate and hydrated pyruvate (2,2-dihydroxypropanoic acid) were also preincubated with the enzyme and NAD+. Kinetic runs were initiated in the resultant solutions at various stages of incubation by the introduction of NADH. The results of the present investigation indicate that hydrated pyruvate is a major inhibitor of lactate dehydrogenase and forms an inhibitory complex with the enzyme and oxidized coenzyme.  相似文献   

16.
《Inorganica chimica acta》1988,149(2):259-264
The bis(N-alkylsalicylaldiminato)nickel(II) complexes Ni(R-sal)2 with R = CH(CH2OH)CH(OH)Ph (I), R = CH(CH3)CH(OH)Ph (II) and R = CH2CH2Ph (III; Ph = phenyl) were prepared and characterized. In the solid state I and II are paramagnetic (μ = 3.2 and 3.3 BM at 20 °C, respectively), whereas III is diamagnetic. It follows from the UV-Vis spectra that in acetone solution I is six-coordinate octahedral and III is four-coordinate planar, the spectrum of II showing characteristics of both modes of coordination. Vis spectrophotometry and stopped-flow spectrophotometry were applied to study the kinetics of ligand substitution in I–III by H2salen (= N,N′-disalicylidene-ethylenediamine) in the solvent acetone at different temperatures. The kinetics follow a second-order rate law, rate = k[H2-salen] [complex]. At 20 °C the sequence of rate constants is k(III):k(II):k(I) = 11 850:40.6:1. The activation parameters are ΔH(I) = 112, ΔH(II) = 40.7, ΔH(III) = 35.7 kJ mol−1 and ΔS(I) = 92, ΔS(II) = −103, ΔS(III) = −89 J K−1 mol−1. The enormous difference in rate between complexes I, II and III, which is less pronounced in methanol, is attributed to the existence of a fast equilibrium planar ⇌ octahedral, which is established in the case of I and II by intramolecular octahedral coordination through the hydroxyl groups present in the organic group R. An A-mechanism is suggested to control the substitution in the sense that the entering ligand attacks the four-coordinate planar complex, the octahedral complex being kinetically inert.  相似文献   

17.
We studied anionic inhibition of the reaction CO2 + OH?? HCO3? catalyzed by human red cell carbonic anhydrase B (I) and C (II), using iodide and cyanate. In the forward reaction with respect to CO2 as the substrate, inhibition was mixed but favoring noncompetitive; the back reaction, with HCO3? as the substrate, yielded strict competitive kinetics. Mean inhibition constants, KI, in the pH range 7.2–7.5 are: iodide, 0.5 mm for enzyme B and 16 mm for C; cyanate, 0.8 μm for B and 20 μm for C. When OH? was considered as the substrate for the forward reaction, cyanate and chloride behaved as competitive inhibitors. The true inhibition constant (KI0) for cyanate (calculated for infinitely low OH?) is 0.4 μm for enzyme B and 4 μm for C. Apart from the difference in anion affinity and some 10-fold higher activity of C > B, the isozymes showed similar patterns of inhibition. Data agree with generally proposed mechanisms describing the active site as ZnH2O with pKa of about 7.  相似文献   

18.
A dye‐sensitized solar cell (DSC) with in situ energy storage capacity is demonstrated using a lead–organohalide electrolyte CH3NH3I·PbCl2 (LOC) to replace the conventional I?/I3? electrolyte. The coupling of lead and iodine in one electrolyte creates a dual‐function rechargeable solar battery that combines the working processes of photoelectrochemical cells with electrochemical batteries. Optimization of the H+ concentration in the electrolyte leads to increased photocharging efficiency and storage. The power conversion efficiency of the LOC–DSC is 8.6% under one sun illumination (AM 1.5, 100 mW cm?2) as a DSC. When operating as a battery, Faraday efficiency can be achieved as high as 81.5% using a bromide‐based CH3NH3Br·PbBr2 (LOB) electrolyte in a DSC configuration. This new cell design suggests a means of combining photovoltaic energy conversion and electrical energy storage.  相似文献   

19.
《Inorganica chimica acta》1988,154(2):183-188
Polymeric complexes of formula [PdCl(TeAr)]n (I) and [Pd(TeAr)2]n (II) are readily obtained by the reaction between Na2[PdCl4] and NaTeAr (ArC6H5, C6H4OCH3−4 and C6H4OCH2CH3−4) in ethanol at room temperature. Chemical and far infrared spectral evidences support alternating chloride and tellurol bridges in I and tellurol bridges in II. While the reaction of I (AtC6H4OCH3−4) with PPh3 in stoichiometric amount results in splitting of chloride bridges and formation of a tellurol bridged dimeric complex [PdCl(TeC6H4OCH3−4)(PPh3)]2 (III), with excess of PPh3, cleavage of both chloride and tellurol bridges leads to the formation of a monomeric compound [PdCl(TeC6H4OCH3−4)(PPh3)2] (IV). Furthermore, the reaction of I (Ar C6H4OCH2CH3−4) with 1,2-bis(diphenyl phosphino)ethane in equimolar ratio also resulted in a monomeric compound [(PdCl(TeC6H4OCH2CH3−4)(diphos)] (V). The complex III (ArC6H4OCH2CH3−4) is also prepared by the reaction between Pd(PPh3)2Cl2 and Ph3SnTeC6H4OCH2CH3−4 in 1:1 molar ratio or between Pd2Cl4(PPh3)2 and Ph3SnTeC6H4OCH2CH3−4 in 1:2 molar ratio in benzene at room temperature. Sodium tetrachloropalladate reacts readily with diarylditellurides in ethanol at 0 °C to form dimeric complexes [PdCl2(ArTeTeAr)]2 (VI). However, at 40 °C or above the same ditellurides form polymeric complexes I with Na2[PdCl4] in ethanol. The complex VI is also obtained by the reaction of Pd(PhCN)2Cl2 with Te2Ar2 in benzene at room temperature. The complexes were characterized by elemental analysis, IR, Raman and 1H NMR spectra and, where possible, by conductivity measurements and molecular weight determinations.  相似文献   

20.
The reaction between [(R-DAB)Rh(PR3)2]+ and molecular hydrogen produces cationic cis-dihydride complexes of Rh(III), of general formula [RhH2(R-DAB)(PR3)2]X. They are stable in air, 1:1 conductors and have been characterized by 1H NMR, 31P NMR, IR and elemental analysis. The tertiary phosphines employed were: PPh3, P(p-C6H4F)3, PMePh2, PEt3, and the R-DAB ligands (RN:CR′CR′:NR)1, Ph-DAB, c-Hex-DAB, NH2-DAB(CH3,CH3), t-but-DAB.The structure of [RhH2(c-Hex-DAB){P(p-C6H4F)3}2]ClO4 has been determined by an X-ray diffraction study. Crystals are orthorhombic, space group Pbnm. Unit cell parameters are: a = 13.032(1), b = 18.166(2), c = 21.449(2) Å, Z = 4, R = 0.081, Rw = 0.082 for 2906 reflections, with I> 3σ(I) the rhodium atom is octahedrally coordinated with the two hydride-hydrogens and c-Hex-DAB in the equatorial plane; the two phosphine ligands are in an axial position bent towards the hydrogens making an angle of 164.9(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