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 421 毫秒
1.
A series of tri- and tetraorganotin compounds containing the optically active 2-(4-isopropyl-2-oxazolinyl)-5-phenyl ligand and tert-butyl, methyl and/or phenyl groups on the tin has been synthesized. All the novel compounds have been characterized, especially by means of the multinuclear NMR investigation, the results of which are discussed. The tin halides, as pairs of diastereoisomers in solution, crystallize in the form of one diastereoisomer. The single-crystal X-ray analysis of tin iodide 10a revealed pseudo-equatorial position of the tert-butyl group opposite to the isopropyl group. In the corresponding diastereomeric tin hydrides values of 1J(1H-117/119Sn) differ significantly, suggesting a different pseudo-axial/equatorial position of the hydrogen atom.  相似文献   

2.
This paper reports the results obtained in a study on the radical hydrostannation of mono- and disubstituted alkynes with bulky triorganotin hydrides using triethylborane as initiator. The addition of trineophyl- (1), tris[(phenyldimethylsilyl)methyl]- (2), and 9-tripticyldimethyltin (3) hydride to eight alkynes was carried out at room temperature leading to vinylstannanes in good to excellent yields and, mostly, with complete stereoselectivity. The results obtained in a study on the relative reactivity of trineophyl- (1), tris[(phenyldimethylsilyl)methyl]- (2), 9-triptycyldimethyltin (3) hydrides, and tri-n-butyltin hydride (29) using the radical reactions between these hydrides and 6-bromo-1-hexene (28) are also reported. Full 1H-, 13C-, and 119Sn NMR characteristics are included.  相似文献   

3.
A series of novel arylgermanium hydrides ArnGeH4–n (n = 1–3) and diaryl(chloro)germanium hydrides Ar2Ge(Cl)H were synthesized and characterized. Systematic preparation and purification were achieved via the lithium chloride–triflic acid and the optimized Grignard route. Arylgermanium hydrides ArnGeH4–n (Ar = 2,5-Me2C6H3, n = 1–3) were characterized by 1H and 73Ge NMR spectroscopy and single crystal X-ray diffractometry.  相似文献   

4.
Hydrogen sorption by several non-stoichiometric ZrMn2-based alloys was studied at pressures up to 50 atm and over a temperature range from 23 to about 200°C. The dissociation pressure of the hydrides is raised by a factor of 500–1000 for ZrMn2T0.8 or ZrMn2T1.2 (T ≡; transition element or Cu) as the host material compared with that for ZrMn2 as the host material. Among the hydrides studied, ZrMn2Co0.8-H exhibited the highest value for the plateau pressure. Measurements of the experimental densities of the non-stoichiometric host materials show good agreement with the substitutional model in which manganese and/or T partially replace zirconium at the zirconium sites. The hydrides have remarkably low heats of formation and entropies of 12–19 kJ (mol H2)−1 and 50–80 J (mol H2)−1 K−1 respectively. The hydrogen absorption or desorption is extremely rapid, e.g. 90% of the hydrogen was released or absorbed in about 1 min. The hydrides studied exhibit features which strongly suggest that they have technological potential.  相似文献   

5.
The N-coordinated tin hydrides containing the chiral 2-(4-isopropyl-2-oxazolinyl)-5-phenyl ligand in the presence of catalytic amounts of tetrakis(triphenylphosphine)palladium gave the corresponding distannanes in good yields. The distannanes have been fully characterized by means of the 1H, 13C, 15N and 117Sn NMR measurements. The J(15N-117/119Sn), J(117Sn-119Sn) couplings and single-crystal X-ray analysis of distannane 3 revealed a tendency towards penta-coordination at the tin center as a result of the Sn-N interaction.  相似文献   

6.
The effective thermal conductivity Ke of Mg2NiH4 and MmNi4FeH5.2 (Mm ≡; misch metal) was measured under steady state conditions as a function of the hydrogen pressure and the temperature. The effective thermal conductivity of the two hydrides increases significantly as a function of the hydrogen pressure up to approximately 40 atm and then remains almost constant. The Ke values for Mg2NiH4 and MmNi4FeH5.2 are 0.83 W m−1 K−1 and 1.05 W m−1 K−1 at 373 K and 273 K respectively and a hydrogen pressure of 40 atm. The results were analysed using the Yagi-Kunii model for the effective thermal conductivity in powder-fluid beds. The solid thermal conductivity Ks of the above two hydrides was then estimated. The most important and significant result of the present work is the conclusion that Ke “saturates” at relatively low values of Ks. For example, the maximum Ke values (above the hydrogen breakaway pressure) at 300 K and a typical hydride void fraction (ϵ = 0.5) change from about 0.4 to about 1.5 W m−1 K−1 for a corresponding Ks change between 1 and 500 W m−1 K−1. A survey of the existing data in the literature for the effective thermal conductivities of powdered metal hydrides supports the above conclusion. The practical meaning of this result is that, in engineering applications of hydrides in which a knowledge of Ke is necessary, it can safely be assumed that Ke ≈ 1–2 Wm−1K−1.  相似文献   

7.
The reaction between basic [(PCP)Pd(H)] (PCP = 2,6-(CH2P(t-C4H9)2)2C6H4) and acidic [LWH(CO)3] (L = Cp (1a), Tp (1b); Cp = η5-cyclopentadienyl, Tp = κ3-hydridotris(pyrazolyl)borate) leads to the formation of bimolecular complexes [LW(CO)2(μ-CO)⋯Pd(PCP)] (4a, 4b), which catalyze amine-borane (Me2NHBH3, tBuNH2BH3) dehydrogenation. The combination of variable-temperature (1H, 31P{1H}, 11B NMR and IR) spectroscopies and computational (ωB97XD/def2-TZVP) studies reveal the formation of an η1-borane complex [(PCP)Pd(Me2NHBH3)]+[LW(CO3)] (5) in the first step, where a BH bond strongly binds palladium and an amine group is hydrogen-bonded to tungsten. The subsequent intracomplex proton transfer is the rate-determining step, followed by an almost barrierless hydride transfer. Bimetallic species 4 are easily regenerated through hydrogen evolution in the reaction between two hydrides.

Bimetallic complexes [LW(CO)2(μ-CO)⋯Pd(PCP)] cooperatively activate amine-boranes for their dehydrogenation via N–H proton tunneling at RDS and H2 evolution from two neutral hydrides.  相似文献   

8.
Triorganotin(IV) hydrides and cyclopentadienides as well as hexaorganodistannanes containing the moiety LCN (2-(N,N-dimethylaminomethyl)phenyl-) as chelating ligand and phenyl, n-butyl or t-butyl substituents were prepared and characterized by NMR and XRD. The compounds reveal trigonal bipyramidal geometry around the central tin atom except for the distannanes in which the tin atom has tetrahedral configuration. The di-n-butyl distannane cannot be oxidized by oxygen or heavier chalcogens and give no tin radical when irradiated by UV light or treated with the TEMPO - free radical at room temperature. LCN(t-Bu)2SnH undergoes reaction in solution toward the corresponding distannane. The hydrostannation reaction of LCN(n-Bu)2SnH with ferrocenylacetylene was investigated. The CO2 activation by LCN(n-Bu)2SnH was also examined.  相似文献   

9.
Cationic platinum(II) hydrides [HPtL2L′]+ (L = phosphines) have been prepared and the ligand steric effects studied; the crystal and molecular structure of trans-[PtH(PCy3)2PPh3] PF6 is described.  相似文献   

10.
A series of nucleophiles was reacted with 1,1,2,4,4,5,7,7,8,8,9,9,9-tridecafluoro-5-trifluoromethyl-3,6-dioxanon-1-ene (1) as a representative of perfluoro(alkyl vinyl ethers). All reactions were completely regioselective with the nucleophilic attack at the terminal carbon atom. Reactions of hydroxy compounds, thiols and sec-amines afforded addition products, but butyllithium, tributylphosphane or complex hydrides caused displacement of vinylic fluorine: butyllithium afforded cis-derivative, while reactions with hydrides and the phosphane led to mixtures of cis- and trans-derivatives. Diethylamine and piperidine adducts displayed the property to substitute hydroxyl for fluorine in hexadecan-1-ol. Molecular properties of hexafluoropropene and perfluoro(methyl vinyl ether) were calculated by ab initio method at the MP2/6-311G(d,p) level of theory and their impact on relative reactivity was estimated.  相似文献   

11.
The model reactions of neopentane hydrogenolysis and hydroisomerization on transition metal hydrides (≡ Si?O)3MIVH (1), (=eqSi?O)2MIVH2 (2), and (≡Si?O)2MIIIH (3) (M = Zr, Ti) immobilized on the surface of silica were studied by density functional theory. It was shown that the hydrogenolysis of neopentane could occur on all the three types of metal hydrides, the catalytic activity of reaction centers increasing along the series M(IV) monohydrides 1 < M(IV) dihydrides 2 < M(III) hydrides 3. At the same time, the isomerization of the hydrocarbon skeleton of neopentane observed experimentally on titanium-based systems could only be explained by the participation of Ti(III) hydrides.  相似文献   

12.
The bonding in closo-boron hydrides, BnH2?n (where n = 5–12), is considered to be the sum of all possible n(n ? 1)/2 boron-boron interactions. The bonding energy u between any pair of boron atoms is taken to depend only upon the internuclear distance d, and to be given by u = 1/d2 ? 1/d. This scheme can be used to rationalize details of molecular geometry and also allows the relative importance of various intramolecular rearrangement pathways to be assessed. This method is compared with alternative approaches to treating the bonding in these molecules.  相似文献   

13.
Alkyl- and aryl-isothiocyanates undergo insertion reactions with platinum metal hydrides to yield the corresponding N-alkyl- and N-aryl-thioformamido
derivatives, the structure and stereochemistry of which have been deduced using 1H NMR data.  相似文献   

14.
The preparation, IR and NMR spectra of 123 platinum hydrides of the general formula, trans-PtHX(PBz3)2 or trans-PtHL(PBz3)2BPh4 (X = a uninegative anionic ligand, L = a neutral donor molecule, Bz = benzyl), are described. Neutral platinum hydrides have been synthesized by the reduction of trans-PtCl2-(PBz3)2 with NaBH4, by the Michaelis—Arbuzov rearrangement, or by metathesis. Cationic hydridoplatinum(II) complexes are obtained from the reaction of trans-PtHX(PBz3)2 (X = Cl or NO3) with a donor molecule (L) in the presence of NaBPh4, or by coordinating a donor molecule through use of PtH(PBz3)2BPh4 · 12CH2Cl2. The observed trends in ν(PtH), τ(H), 1J(PtH) and 1J(PtP) in a series of the hydridobenzylphosphineplatinum(II) complexes are discussed in terms of “trans- or cis-influences”, defined as the ability of a ligand to weaken the bond trans or cis to itself. The data support the view that a donor atom trans to the hydridic ligand is important in determining the strength of the PtH bond in this series. Some remarks on the distinctive characteristics of some complexes, e.g., dissociation of coordinated cycloalkanone from platinum(II) or stereochemical non-rigidity of the sym-dimethylurea ligand, are included. Tricyclohexylphosphine analogs also have been prepared for comparison.  相似文献   

15.
Iron and ruthenium classical and non-classical hydrides of the type [MH(N–N)P3]+ and [M(η2-H2)(N–N)P3]2+ {M = Fe, Ru; N–N = 2,2′-bipyridine (bpy), 1,10-phenanthroline (phen); P = phosphites} were reported in 2004 together with an evaluation of the pseudo-aqueous pKa values of the η2-H2 complexes. The non-classical hydrides, even if doubly charged, showed a relatively low acidity, their pKa values ranging between −5.4 and −4.3. Moreover, ruthenium(II) derivatives showed to be more acidic than the corresponding iron(II) complexes. Information about the structural and electronic proprieties of complexes of this type, which allowed to better understand the role of both the metal centres and the ancillary ligands in the acidity of the co-ordinated hydrogen molecule, was obtained on the basis of DFT B3LYP calculations.  相似文献   

16.
Integral enthalpies of absorption and desorption of hydrogen by hyperstoichiometric ZrMn2T0.8 (T = Mn, Fe, Co, Ni, and Cu) and stoichiometric ZrMn2- and ZrCr2-based alloys have been determined. The measured enthalpies range from ~24 to ~41 kJ/mole H2. The ΔH values for hydrides formed by the series of metallic hosts ZrMn2T0.8 are smaller than that for ZrMn2, accounting for the enhanced dissociation pressures of the ZrMn2T0.8 hydrides. In the series of ZrMn2T0.8 hydrides there is a pronounced minimum for hydride of ZrMn2Co0.8, accounting for the extraordinarily high decomposition pressure of this system. Site occupancies, provided by published neutron diffraction studies, were used to calculate configurational entropies of ZrCr2 hydrides and related systems. Results obtained were in fair agreement with experiment.  相似文献   

17.
Significantly fluorinated triarylmethyl cations have long attracted attention as potentially accessible highly reactive carbocations, but their isolation in a convenient form has proved elusive. We show that abstraction of chloride with a cationic silylium reagent leads to the facile formation of di-, tetra-, and hexafluorinated trityl cations, which could be isolated as analytically pure salts with the [HCB11Cl11] counterion and are compatible with (halo)arene solvents. The F6Tr+ cation carrying six meta-F substituents was computationally predicted to possess up to 20% higher hydride affinity than the parent triphenylmethyl cation Tr+. We report that indeed F6Tr+ displays reactivity unmatched by Tr+. F6Tr+ at ambient temperature abstracts hydrides from the C–H bonds in tetraethylsilane, mesitylene, methylcyclohexane, and catalyzes Friedel–Crafts alkylation of arenes with ethylene, while Tr+ does none of these.

Hexakis(meta-fluoro)substituted triphenylmethyl cation (F6Tr+) was isolated with a carborane partner anion. Its significantly higher hydride affinity compared to the parent (Tr+) allows it to abstract hydrides from a variety of C(sp3)–H bonds.  相似文献   

18.
The complexes [(C5Me5)Ir(η6-arene)][BF4]2 (arene = toluene, toluene-d8, t-butylbenzene, methoxybenzene, chlorobenzene, o-xylene, p-xylene, tetralin and phenol) were prepared from the arene and reduced with NaBH4 to the η5-cyclohexadienyl complexes. Attack was exo at the arene and, with one exception, never at the substituent. Toluene showed no site preference but t-butylbenzene was attacked preferentially para, and chlorobenzene, ortho. Methoxybenzene was attacked ipso as well as ortho, meta (predominant), and para, and phenol gave only the meta-isomer. p-Xylene gave one isomer and o-xylene and tetralin gave two. Further reduction occurred on reaction with stronger hydride reducers (e.g., sodium bis(methoxyethoxy)dihydroaluminate) to give mixtures of 1- and 2-substituted cyclohexa-1,3-diene complexes (t-Bu, 2- ( > 95%); Me, 1- (25%), 2- (75%); Cl, 1- ( > 95%); and OMe, 1- (33%), 2- (67%)). The p-xylene complex gave a mixture of the η4-1,4-dimethylcyclohexa-1,3- and 1,4-diene complexes. Reaction of the cyclohexadiene complexes with HCl gas gave the free substituted cyclohexenes and [(C5Me5Ir)2Cl4]. The product from t-butylbenzene was predominantly (92%) the 3-substituted cyclohexene; that isomer (65%) and the 1-isomer (34%) were formed from toluene and the 1- (34%) and the 4-isomer (58%) were formed from chlorobenzene. Phenol gave only cyclohexanone. Overall these reactions yield the cyclohexene from the substituted benzene by addition of two hydrides and two protons and the iridium can be recycled.  相似文献   

19.
The ground state Rydberg—Klein—Rees (RKR) potentials and the corresponding molecular constants of the alkali hydrides, recommended in a recent article by Stwalley et al. (J. Phys. Chem. Ref. Data, to be published [1]) are critically evaluated in the framework of the reduced potential curve (RPC) scheme. A comparison with the older RPC analysis of the ground states of the alkali hydrides is briefly discussed. The efficiency of the RPC method for the detection of errors in the RKR potential (spectroscopic constants) and for the estimation of the dissociation energy is emphasized. Although the RKR potentials of NaH and RbH are known only up to 54 and 57% of De, respectively, the RPC method permitted here at least a substantial reduction of the uncertainty in the lower limit of De(NaH) (by 70 cm−1) and in the lower and upper limits of De(RbH) (by 250 and 500 cm−1, respectively) which are now estimated as 15 870, 14 230 and 14 680 cm−1, respectively. The RPC picture even suggests that the values 14 380 and 14 580 cm−1 may possibly be taken as reasonable limits for De(RbH). Accurate extensions of the inner wings of the potentials of NaH, RbH and CsH were calculated using the generalized reduced potential curve (GRPC) method. The limit of error of these extensions should be smaller than 0.002 Å if the potentials are correct.  相似文献   

20.
The reaction of the donor-stabilized silylene complex cis-Cp1(CO)2(H)WSiHPh · THF (3, Cp1 = η5-C5Me5) with LiAlH4 followed by the protonation of the resulting Li[Cp1(CO)2W(H)(SiH2Ph)] (4) with excess CF3COOH afforded the trihydride complex Cp1(CO)2WH3 (6). The structure of 6 was characterized using variable-temperature NMR studies and X-ray crystal analysis. Deprotonation of 6 with KH gave the anionic dihydride complex K[Cp1(CO)2WH2] (7), which was converted into the dichlorosilyl dihydride complex Cp1(CO)2W(H)2(SiHCl2) (8) on treatment with trichlorosilane. The X-ray crystal analysis of 8 revealed that it adopts a distorted pseudo-octahedral structure with a short W–Si bond, long Si–Cl bonds, and short contacts between the hydrides and silicon atom. Along with these structural features, the conformation of the silyl ligand around the W–Si bond may suggest the presence of a double interligand hypervalent interaction between the dichlorosilyl and hydrides ligands.  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