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.
Metal hydrides are of fundamental importance in chemistry, both as solid‐state materials and molecular compounds. The first low‐valent molecular metal cluster containing an interstitial four‐coordinate hydride in a tetrahedral site is decribed, which undergoes hydride migration from the surface to the tetrahedral cavity of the cluster upon coordination of a [AuPPh3]+ fragment. The [HFe4(CO)12(AuPPh3)2]? mono‐anion, which contains a surface μ3‐H, was obtained from the reaction of [HFe4(CO)12]3? with two equivalents of [Au(PPh3)Cl]. This is, in turn, transformed into the neutral [HFe4(CO)12(AuPPh3)3] upon addition of a third [AuPPh3]+ fragment, with concomitant migration of the unique hydride from the surface of the cluster to its tetrahedral cavity. All of these species have been fully characterized in solution by means of IR and multinuclear NMR spectroscopy, and in the solid state by single‐crystal X‐ray diffractometry.  相似文献   

2.
3.
Photolysis of a hexane solution containing 1,1′- bis (trimethylsilylethynyl)ferrocene ( 1 ) and Fe (CO)5, under argon at 0 °C led to the formation of dinuclear complexes [Fe (CO)222 – C (SiMe3) = C(C5H4)FeC(C5H4) = C (SiMe3)Fe (CO)3}–μ–CO] ( 2 ) and [Fe (CO)222–C (SiMe3) = C(C5H5)–C(C5H5) = C (SiMe3)Fe (CO)3}–μ–CO] ( 3 ). DFT calculations support the experimentally observed demetalation of ferrocene unit of 2 to 3 in presence of water. These compounds were comprehensively characterized by IR and 1H and 13C NMR spectroscopy and crystallographically ( 1 and 3 ).  相似文献   

4.
5.
6.
The dehydration of primary amides to their corresponding nitriles using four [PSiP]-pincer hydrido iron complexes 1–4 [(2-Ph2PC6H4)2MeSiFe(H)(PMe3)2 ( 1 ), (2-Ph2PC6H4)2HSiFe(H)(PMe3)2 ( 2 ), (2-(iPr)2PC6H4)2HSiFe(H)(PMe3)2 ( 3 ) and (2-(iPr)2PC6H4)2MeSiFe(H)(PMe3)2 ( 4 )] as catalysts in the presence of (EtO)3SiH as dehydrating reagent was explored in the good to excellent yields. It was proved for the first time that Lewis acid could significantly promote this catalytic system under milder reaction conditions than other Lewis acid-promoted system, such as shorter reaction time or lower reaction temperature. This is also the first example that dehydration of primary amides to nitriles was catalyzed by silyl hydrido iron complexes bearing [PSiP]-pincer ligands with Lewis acid as additive. This catalytic system has good tolerance for many substituents. Among the four iron hydrides 1 is the best catalyst. The effects of substituents of the [PSiP]-pincer ligands on the catalytic activity of the iron hydrides were discussed. A catalytic reaction mechanism was proposed. Complex 4 is a new iron complex and was fully characterized. The molecular structure of 4 was determined by single crystal X-ray diffraction.  相似文献   

7.
8.
9.
10.
11.
12.
The reaction of Fe3(CO)12 ( 13 ) with 7,8‐dithiabicyclo[4.2.1]nona‐2,4‐diene 7‐exo‐oxide ( 12 ) yields the sulfenato‐thiolato complex 14 , which is used as starting material for further reactions. The disulfenato complex 17 is obtained by using one equivalent of dimethyldioxirane (DMD), and the monoepoxide 18 is prepared by the oxidation of 14 with an excess of DMD. Complex 14 can be converted to the monophosphine complexes 19a and 19b by subsequent substitution of one CO ligand using trimethylaminoxide Me3NO and triphenylphosphine PPh3. Additional substitution reactions are done with 17 by using acetonitrile as a ligand to form 20a and 20b . In the electrochemical part of the paper, the reactions of the reduced iron species 14 , 15 , 17 , and 19a are studied.  相似文献   

13.
Two different procedures based on lead hydride generation for determination of lead in commercial iron oxide pigments have been evaluated. As the procedure based on the prior acid-dissolution of the samples to give a 1M HCl final medium led to a high relative standard deviation (6.5%) an alternative procedure based on the use of slurries was studied. The samples were suspended in water containing 0.01% hexametaphosphate, and lead hydride was generated from a 0.7M nitric acid and 14% ammonium peroxodisulphate medium by addition of 10% tetrahydroborate solution. In this way, an improvement in reproducibility and sensitivity as well as a saving of time and effort was achieved. The procedure based on the use of a suspension of the samples is therefore recommended.  相似文献   

14.
The reactions of nitro and azo compounds with carbon monoxide were studied in the presence of iron carbonyl catalysts. It was shown that these catalytic systems differ substantially from Pd- and Rh-containing catalysts. In the case of the iron catalysts, the products of coupling of molecules are formed as intermediates and azo compounds are the final reaction products. The reactions involving the palladium and rhodium catalysts proceed without the intermediate formation of the coupling products and lead to isocyanates or carbamates. When combined using PdCl2 and Fe(CO)5/Al2O3, the catalysts inhibit each other, especially in the presence of pyridine.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 2460–2463, October, 1996.  相似文献   

15.
The synthesis, characterization and properties of [PdH(dppe)(2)](+)CF(3)SO(3) (-).0.125 THF (1; dppe=1,2-bis(diphenylphosphanyl)ethane) and its SbF(6) (-) (1') and BF(4) (-) (1") analogues, the missing members of the [MH(dppe)(2)](+)X(-) (M=Ni, Pd, Pt) family, are described. The Pd hydrides are not stable in solution and can react as proton or hydride donors with formation of dihydrogen, [Pd(dppe)(2)](2+) and [Pd(dppe)(2)]. Complexes 1-1" react with carbocations and carbanions by transferring a hydride and a proton, respectively. Such H(-) or H(+) transfer occurs also towards unsaturated compounds, for example, hydrogenation of a C=C double bond. Accordingly, 1 can hydrogenate methyl acrylate to methyl propionate. Complex 1" is an effective (hourly turnover frequency=16) and very selective (100 %) catalyst for the hydrogenation of cyclohexen-2-one to cyclohexanone with dihydrogen under mild conditions. Density functional calculations coupled with a dielectric continuum model were carried out to compute the energetics of the hydride/proton transfer reactions, which were used to rationalize some of the experimental findings. Theory provides strong support for the thermodynamic and kinetic viability of a tetracoordinate Pd complex as an intermediate in the reactions.  相似文献   

16.
The reactions of [Ru(N2)(PR3)(‘N2Me2S2’)] [‘N2Me2S2’=1,2‐ethanediamine‐N,N′‐dimethyl‐N,N′‐bis(2‐benzenethiolate)(2?)] [ 1 a (R=iPr), 1 b (R=Cy)] and [μ‐N2{Ru(N2)(PiPr3)(‘N2Me2S2’)}2] ( 1 c ) with H2, NaBH4, and NBu4BH4, intended to reduce the N2 ligands, led to substitution of N2 and formation of the new complexes [Ru(H2)(PR3)(‘N2Me2S2’)] [ 2 a (R=iPr), 2 b (R=Cy)], [Ru(BH3)(PR3)(‘N2Me2S2’)] [ 3 a (R=iPr), 3 b (R=Cy)], and [Ru(H)(PR3)(‘N2Me2S2’)]? [ 4 a (R=iPr), 4 b (R=Cy)]. The BH3 and hydride complexes 3 a , 3 b , 4 a , and 4 b were obtained subsequently by rational synthesis from 1 a or 1 b and BH3?THF or LiBEt3H. The primary step in all reactions probably is the dissociation of N2 from the N2 complexes to give coordinatively unsaturated [Ru(PR3)(‘N2Me2S2’)] fragments that add H2, BH4?, BH3, or H?. All complexes were completely characterized by elemental analysis and common spectroscopic methods. The molecular structures of [Ru(H2)(PR3)(‘N2Me2S2’)] [ 2 a (R=iPr), 2 b (R=Cy)], [Ru(BH3)(PiPr3)(‘N2Me2S2’)] ( 3 a ), [Li(THF)2][Ru(H)(PiPr3)(‘N2Me2S2’)] ([Li(THF)2]‐ 4 a ), and NBu4[Ru(H)(PCy3)(‘N2Me2S2’)] (NBu4‐ 4 b ) were determined by X‐ray crystal structure analysis. Measurements of the NMR relaxation time T1 corroborated the η2 bonding mode of the H2 ligands in 2 a (T1=35 ms) and 2 b (T1=21 ms). The H,D coupling constants of the analogous HD complexes HD‐ 2 a (1J(H,D)=26.0 Hz) and HD‐ 2 b (1J(H,D)=25.9 Hz) enabled calculation of the H? D distances, which agreed with the values found by X‐ray crystal structure analysis ( 2 a : 92 pm (X‐ray) versus 98 pm (calculated), 2 b : 99 versus 98 pm). The BH3 entities in 3 a and 3 b bind to one thiolate donor of the [Ru(PR3)(‘N2Me2S2’)] fragment and through a B‐H‐Ru bond to the Ru center. The hydride complex anions 4 a and 4 b are extremely Brønsted basic and are instantanously protonated to give the η2‐H2 complexes 2 a and 2 b .  相似文献   

17.
18.
The reaction of [Co(2)(CO)(8)] with (CF(3))(3)BCO in hexane leads to the Lewis acid-base adduct [Co(2)(CO)(7)CO--B(CF(3))(3)] in high yield. When the reaction is performed in anhydrous HF solution [Co(CO)(5)][(CF(3))(3)BF] is isolated. The product contains the first example of a homoleptic metal pentacarbonyl cation with 18 valence electrons and a trigonal-bipyramidal structure. Treatment of [Co(2)(CO)(8)] or [Co(CO)(3)NO] with NO(+) salts of weakly coordinating anions results in mixed crystals containing the [Co(CO)(5)](+)/[Co(CO)(2)(NO)(2)](+) ions or pure novel [Co(CO)(2)(NO)(2)](+) salts, respectively. This is a promising route to other new metal carbonyl nitrosyl cations or even homoleptic metal nitrosyl cations. All compounds were characterized by vibrational spectroscopy and by single-crystal X-ray diffraction.  相似文献   

19.
The equilibrium structures and vibrational frequencies of the iron complexes [Fe0(CN)n(CO)5?n]n? and [FeII(CN)n(CO)5?n]2?n (n = 0–5) have been calculated at the BP86 level of theory. The Fe0 complexes adopt trigonal bipyramidal structures with the cyano ligands occupying the axial positions, whereas corresponding Fe2+ complexes adopt square pyramidal structures with the cyano ligands in the equatorial positions. The calculated geometries and vibrational frequencies of the mixed iron Fe0 carbonyl cyanide complexes are in a very good agreement with the available experimental data. The nature of the Fe? CN and Fe? CO bonds has been analyzed with both charge decomposition and energy partitioning analysis. The results of energy partitioning analysis of the Fe? CO bonds shows that the binding interactions in Fe0 complexes have 50–55% electrostatic and 45–50% covalent character, whereas in Fe2+ 45–50% electrostatic and 50–55% covalent character. There is a significant contribution of the π‐ orbital interaction to the Fe? CO covalent bonding which increases as the number of the cyano groups increases, and the complexes become more negatively charged. This contribution decreases in going from Fe0 to Fe2+ complexes. Also, this contribution correlates very well with the C? O stretching frequencies. The Fe? CN bonds have much less π‐character (12–30%) than the Fe? CO bonds. © 2010 Wiley Periodicals, Inc. J Comput Chem, 2010  相似文献   

20.
The iron dithiolene compounds [Fe2(mnt)4]2− [1]2− and [Fe(NO)(mnt)2]n (n = 1−, [2]1−; n = 2−, [2]2−) ([mnt]2− = maleonitriledithiolate = [(NC)2C2S2]2−) have been characterized structurally by X-ray diffraction as their [Et4N]+ salts at 100 K. Dianion [2]2− is prepared from [2]1− by reduction with Na[Et3BH] and is observed to have a bent Fe-NO angle at 149.9(5)° in contrast to the linear configuration of Fe-NO in [2]1− (180.0°). The change from linear to bent binding mode for NO, an increase of more than 0.1 Å in the Fe-N bond length, and the relative invariance of the Fe-S distances for [2]2− versus [2]1− indicate that the NO ligand is the site of reduction. The [Et3NH]+ complex of [2]1− was also identified by crystallography and found to have hydrogen bonding contacts between [Et3NH]+ and the cyano nitrogen atom of an [mnt]2− ligand. Furthermore, relatively close S?S contacts (3.602-3.615 Å) occur between [2]1− anions, which pack together in an offset, head-to-head fashion. These S?S contacts are absent in the structure of [Et4N][2]. Infrared spectra show an energy decrease for, and a significant broadening of, the NO bond stretching absorption peak in [2]2−, which is consistent with a bent NO ligand sampling a range of conformations both by facile pivoting about the Fe-N axis and by a breathing of the Fe-NO angle.  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