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 46 毫秒
1.
Surface organometallic chemistry (SOMC) on silica materials is a prominent approach for the generation of highly active heterogenized polymerization catalysts. Despite advanced methods of characterization, the elucidation of the catalytically active surface species remains a challenging task. Alkylated rare‐earth metal siloxide complexes can be regarded as molecular models of respective covalently bonded alkylated surface species, primarily used for 1,3‐diene polymerization. Here, we performed both salt metathesis reactions of [Y(MMe4)3] (M = Al, Ga) with [K{OSi(OtBu)3}] and alkylation reactions of [Y{OSi(OtBu)3}3]2 with AlMe3. The obtained complexes [Y(CH3)[(AlMe2){OSi(OtBu)3}2](AlMe4)]2, [Y(CH3)[(AlMe2){OSi(OtBu)3}2]‐{OSi(OtBu)3}], [Y{OSi(OtBu)3}3(μ‐Me)Y(μ‐Me)2Y{OSi(OtBu)3}2(AlMe4)], and [Y(CH3)(GaMe4){OSi(OtBu)3}]2 represent rare examples of organoyttrium species with terminal methyl groups. The formation and purity of the mixed methyl/siloxy yttrium complexes could be enhanced by treating [Y(MMe4)3] with [K(MMe2){OSi(OtBu)3}2]n (M=Al: n=2; M=Ga: n=∞). Complexes [K(MMe2){OSi(OtBu)3}2]n were obtained by addition of [K{OSi(OtBu)3}] to [Me2M{OSi(OtBu)3}]2. Deeper insight into the fluxional behavior of the mixed methyl/siloxy yttrium complexes in solution was gained by 1H and 13C NMR spectroscopic studies at variable temperature and 1H–89Y HSQC NMR spectroscopy.  相似文献   

2.
Thermolysis of the nitride‐bridged diuranium(IV) complex Cs{(μ‐N)[U(OSi(OtBu)3)3]2} ( 1 ) showed that the bridging nitride behaves as a strong nucleophile, promoting N?C bond formation by siloxide ligand fragmentation to yield an imido‐bridged siloxide/silanediolate diuranium(IV) complex, Cs{(μ‐NtBu)(μ‐O2Si(OtBu)2)U2(OSi(OtBu)3)5}. Complex 1 displayed reactivity towards CS2 and CO2 at room temperature that is unprecedented in f‐element chemistry, affording diverse N‐functionalized products depending on the reaction stoichiometry. The reaction of 1 with two equivalents of CS2 yielded the thiocyanate/thiocarbonate complex Cs{(μ‐NCS)(μ‐CS3)[U(OSi(OtBu)3)3]2} via a putative NCS?/S2? intermediate. The reaction of 1 with one equivalent of CO2 resulted in deoxygenation and N?C bond formation, yielding the cyanate/oxo complex Cs{(μ‐NCO)(μ‐O)[U(OSi(OtBu)3)3]2}. Addition of excess CO2 to 1 led to the unprecedented dicarbamate product Cs{(μ‐NC2O4)[U(OSi(OtBu)3)3]2}.  相似文献   

3.
Syntheses and Reactions of Aluminium Alkoxide Compounds Al(OcHex)3 ( 1 ) can be synthesized by the reaction of Al with cyclohexanol under evolving of H2 in boiling xylene. [Li{Al(OCH2Ph)4}] ( 2 ) was obtained by treatment of PhCH2OH with a 1 M solution of LiAlH4 in THF. [{(THF)Li}2{Al(OtBu)4}Cl] ( 3 ) is the result of the reaction of four equivalents of LiOtBu on AlCl3 in THF. 3 is the educt for the reactions with the Lewis‐acids InCl3 and FeCl3 in THF leading to the metalates [{(THF)2Li}2{Al(OtBu)4}] · [MCl4] [M = In ( 4 ), Fe ( 5 )]. The attempt to react InCl3 with four equivalents of LiOtBu leads to only one isolated and characterized product, the complex [Li4(OtBu)3(THF)3Cl]2 · THF ( 6 · THF), which can also be synthesized by the treatment of LiCl with three equivalents of LiOtBu in THF. 1–6 · THF were characterized by NMR, IR and MS techniques as well as by X‐ray structure determinations. According to them, 1 , which is tetrameric in solution, is the first structurally characterized example of the proposed trimer form of aluminium alkoxides [ROAl{Al(OR)4}2] with a central trigonal bipyramidal coordinated Al atom. 2 forms a coordination polymer with a distorted tetrahedral coordination sphere of Li and Al, running along [100]. The trinuclear structure skeleton [{(THF)2Li}2{Al(OtBu)4}]+ is still present in the isotypical metalates 4 and 5 . The counter ions [MCl4] possess nearly Td symmetry. The remarkable structural motif of 6 · THF are two heterocubanes [Li4(OtBu)3(THF)3Cl] dimerized by Li–Cl bonds.  相似文献   

4.
Two new arene inverted‐sandwich complexes of uranium supported by siloxide ancillary ligands [K{U(OSi(OtBu)3)3}2(μ‐η66‐C7H8)] ( 3 ) and [K2{U(OSi(OtBu)3)3}2(μ‐η66‐C7H8)] ( 4 ) were synthesized by the reduction of the parent arene‐bridged complex [{U(OSi(OtBu)3)3}2(μ‐η66‐C7H8)] ( 2 ) with stoichiometric amounts of KC8 yielding a rare family of inverted‐sandwich complexes in three states of charge. The structural data and computational studies of the electronic structure are in agreement with the presence of high‐valent uranium centers bridged by a reduced tetra‐anionic toluene with the best formulation being UV–(arene4?)–UV, KUIV–(arene4?)–UV, and K2UIV–(arene4?)–UIV for complexes 2 , 3 , and 4 respectively. The potassium cations in complexes 3 and 4 are coordinated to the siloxide ligands both in the solid state and in solution. The addition of KOTf (OTf=triflate) to the neutral compound 2 promotes its disproportionation to yield complexes 3 and 4 (depending on the stoichiometry) and the UIV mononuclear complex [U(OSi(OtBu)3)3(OTf)(thf)2] ( 5 ). This unprecedented reactivity demonstrates the key role of potassium for the stability of these complexes.  相似文献   

5.
Metalat Ions [Al(OR)4] as Chelating Ligands for Transition Metal Cations Waterfree CoCl2 can be reacted with [{Li(Diglyme)}{Al(OtBu)4}] in THF to the complex [Li(THF)4][{CoCl2}{Al(OtBu)4}]. Addition of diglyme to the reaction mixtures gives the blue compound [Li(diglyme)2][{CoCl2}{Al(OtBu)4}] ( 1 ). According to this procedure the FeII complex [Li(Diglyme)2][{FeCl2}2{Al(OtBu)4}] ( 2 ) was formed by treatment of FeCl2 with Li[Al(OtBu)4]. [{Li(diglyme)}{Al(OtBu)4}] in THF/diglyme can be used as alkoxide transfer reagent on TiCl4 to give the neutral complex [TiCl2(OtBu)2(diglyme)] ( 3 ). The sky‐blue salt [Li(THF)4]2[{CoCl2}3{Al(OCH2Ph)4}2] ( 4 ) was obtained by reaction of Li[Al(OCH2Ph)4] with CoCl2 in THF. By treatment of 4 with diglyme ligand redistribution was observed giving the sky‐blue compound [Li(Diglyme)2]2[{CoCl2}3{Al(OCH2Ph)4}2] ( 5 ) and the violet salt [Li(Diglyme)2]2[Co2Cl5(OCH2Ph)] ( 6 ). A similar salt can be synthesized also directly from Li[Al(OtBu)4] and CoCl2 in diglyme to give [Li(Diglyme)2]2[Co2Cl5(OtBu)] ( 7 ). 1 — 7 were characterized by IR spectroscopy, partly by mass spectrometry and X‐ray analyses. UV‐VIS spectra were recorded from 1 and 5 . According to the X‐ray analyses the MII ions as well as the AlIII ions are coordinated distorted tedrahedrally. In 1 , 2 , 4 und 5 the unit [Al(OR)4] acts a chelating ligand as desired.  相似文献   

6.
Reaction of the secocubane [Sn32‐NHtBu)22‐NtBu)(μ3‐NtBu)] ( 1 ) with dibutylmagnesium produces the heterobimetallic cubane [Sn3Mg(μ3‐NtBu)4] ( 4 ) which forms the monochalcogenide complexes of general formula [ESn3Mg(μ3‐NtBu)4] ( 5 a , E=Se; 5 b , E=Te) upon reaction with elemental chalcogens in THF. By contrast, the reaction of the anionic lithiated cubane [Sn3Li(μ3‐NtBu)4]? with the appropriate quantity of selenium or tellurium leads to the sequential chalcogenation of each of the three SnII centres. Pure samples of the mono‐ or dichalcogenides are, however, best obtained by stoichiometric redistribution reactions of [Sn3Li(μ3‐NtBu)4]? and the trichalcogenides [E3Sn3Li(μ3‐NtBu)4]? (E=Se, Te). These reactions are conveniently monitored by using 119Sn NMR spectroscopy. The anion [Sn3Li(μ3‐NtBu)4]? also acts as an effective chalcogen‐transfer reagent in reactions of selenium with the neutral cubane [{Snμ3‐N(dipp)}4] ( 8 ) (dipp=2,6‐diisopropylphenyl) to give the dimer [(thf)Sn{μ‐N(dipp)}2Sn(μ‐Se)2Sn{μ‐N(dipp)}2Sn(thf)] ( 9 ), a transformation that results in cleavage of the Sn4N4 cubane into four‐membered Sn2N2 rings. The X‐ray structures of 4 , 5 a , 5 b , [Sn3Li(thf)(μ3‐NtBu)43‐Se)(μ2‐Li)(thf)]2 ( 6 a ), [TeSn3Li(μ3‐NtBu)4][Li(thf)4] ( 6 b ), [Te2Sn3Li(μ3‐NtBu)4][Li([12]crown‐4)2] ( 7 b′′ ) and 9 are presented. The fluxional behaviour of cubic imidotin chalcogenides and the correlation between NMR coupling constants and tin–chalcogen bond lengths are also discussed.  相似文献   

7.
Titanosilicon mesoporous materials with a high content of titanium (up to Ti/Si 0.36) and high adsorptive characteristics have been obtain from the titanosilicon ester molecular precursors (BuO)2Ti[OSi(O t Bu)3]2 and BuOTi[OSi(O t Bu)3]3 by template synthesis in the presence of micelle forming surfactants. A characteristic of these mesoporous matrices is the considerable size of the walls of the mesopores (2-4 nm) which may results from the use of the bulky titanosilicon building blocks in the template synthesis.  相似文献   

8.
Cleavage of dihydrogen is an important step in the industrial and enzymatic transformation of N2 into ammonia. The reversible cleavage of dihydrogen was achieved under mild conditions (room temperature and 1 atmosphere of H2) by the molecular uranium nitride complex, [Cs{U(OSi(OtBu)3)3}2(μ‐N)] 1, leading to a rare hydride–imide bridged diuranium(IV) complex, [Cs{U(OSi(OtBu)3)3}2(μ‐H)(μ‐NH)], 2 that slowly releases H2 under vacuum. This complex is highly reactive and quickly transfers hydride to acetonitrile and carbon dioxide at room temperature, affording the ketimide‐ and formate‐bridged UIV species [Cs{U(OSi(OtBu)3)3}2(μ‐NH)(μ‐CH3CHN)], 3 and [Cs{U(OSi(OtBu)3)3}2(μ‐HCOO)(μ‐NHCOO)], 4 .  相似文献   

9.
Interaction of ptert‐butylcalix[8]areneH8 (L8H8) with [NaVO(OtBu)4] (formed in situ from VOCl3) afforded the complex [Na(NCMe)5][(VO)2L8H]?4 MeCN ( 1 ?4 MeCN). Increasing [NaVO(OtBu)4] to 4 equiv led to [Na(NCMe)6]2[(Na(VO)4L8)(Na(NCMe))3]2?10 MeCN ( 2 ?10 MeCN). With adventitious oxygen, reaction of 4 equiv of [VO(OtBu)3] with L8H8 afforded the alkali‐metal‐free complex [(VO)4L83‐O)2] ( 3 ); solvates 3 ?3 MeCN and 3 ?3 CH2Cl2 were isolated. For the lithium analogue, the order of addition had to be reversed such that lithium tert‐butoxide was added to L8H8 and then treated with 2 equiv of VOCl3; crystallisation afforded [(VO2)2Li6[L8](thf)2(OtBu)2(Et2O)2]?Et2O ( 4 ?Et2O). Upon extraction into acetonitrile, [Li(NCMe)4][(VO)2L8H]?8 MeCN ( 5 ?8 MeCN) was formed. Use of the imido precursors [V(NtBu)(OtBu)3] and [V(Np‐tolyl)(OtBu)3] and L8H8, afforded [tBuNH3][{V(p‐tolylN)}2L8H]?3 1/2 MeCN ( 6 ?3 1/2 MeCN). The molecular structures of 1 to 6 are reported. Complexes 1 , 3 , and 4 were screened as precatalysts for the polymerisation of ethylene in the presence of cocatalysts at various temperatures and for the copolymerisation of ethylene with propylene. Activities as high as 136 000 g (mmol(V) h)?1 were sometimes achieved; higher molecular weight polymers could be obtained versus the benchmark [VO(OEt)Cl2]. For copolymerisation, incorporation of propylene was 7.1–10.9 mol % (compare 10 mol % for [VO(OEt)Cl2]), although catalytic activities were lower than [VO(OEt)Cl2].  相似文献   

10.
Herein, new complexes containing the [Ph2PCH2S(NtBu)3]? anion are presented, supplying three imido nitrogen atoms and a remote phosphorus atom as potential donor sites to main group and transition‐metal cations. The lithiated complex [(tmeda)Li{(NtBu)3SCH2PPh2}] ( 1 ) is an excellent starting material in transmetalation reactions. Herein, the transition‐metal complexes [M{(NtBu)3SCH2PPh2}2] (M=Mn ( 2 ), Ni ( 3 ), Zn ( 4 )) were synthesized and structurally characterized. Their isotypical molecules show SN2 chelation and no employment of the adjacent phosphorus atom in coordination. The third pendent imido group is always twisted toward the vacant face of the tetrahedrally coordinated sulfur atom.  相似文献   

11.
Alkylation of spiro[fluorene-9,3’-indazole] at N(1) and N(2) with tBuCl affords the nitrenium cations [C6H4N2(tBu)C(C12H8)][BF4], 1 and 2 , respectively. Compound 1 converts to 2 over the temperature range 303–323 K with a free energy barrier of 28±5 kcal mol−1. Reaction of 1 with PMe3 afforded the N-bound phosphine adduct [C6H4N(tBu)N(PMe3)C(C12H8)]BF4] 3 . However, phosphines attack 2 at the para-carbon atom of the aryl group with concurrent cleavage of N(2)−C(1) bond and proton migration to C(1) affording [(R3P)C6H3NN(tBu)CH(C12H8)][BF4] (R=Me 4 , nBu 5 ). Analogous reactions of 1 and 2 with the carbene SIMes prompt attack at the para-carbon with concurrent loss of H. affording the radical cation salts [(SIMes)C6H3N(tBu)NC(C12H8).][BF4] 6 and [(SIMes)C6H3NN(tBu)C(C12H8).][BF4] 7 , whereas reaction of 2 with BAC gives the Lewis acid-base adduct, [C6H4N(BAC)N(tBu)C(C12H8)][BF4] 8 . Finally, reactions of 1 and 2 with KPPh2 result in electron transfer affording (PPh2)2 and the persistent radicals C6H4N(tBu)NC(C12H8). and C6H4NN(tBu)C(C12H8).. The detailed reaction mechanisms are also explored by extensive DFT calculations.  相似文献   

12.
We have investigated the coordination of alkanide and alkynide anions to the coordinatively unsaturated aluminium atoms of the methylene‐bridged dialuminium compound R2Al‐CH2‐AlR2 [ 1 , R = CH(SiMe3)2]. Treatment of 1 with the corresponding lithium derivatives in the presence of a small excess of TMEN (TMEN = tetramethylethylenediamine) yielded mono‐adducts [M]+[R2Al‐CH2‐AlR2R'] [ 2a , M = Li(TMEN)2, R' = Me; 2b , M = Li(TMEN)2, R' = n‐Bu; 3a , M = Li(TMEN)2, R' = C≡C‐SiMe3; 3b , M = Li(TMEN)2, R' = C≡C‐t‐Bu; 3d , M = Li(DME)3, R' = C≡C‐Ph; 3e , M = Li(TMEN)2, R' = C≡C‐PPh2)] and bis‐adducts [Li(TMEN)2]+[LiCH2(AlR2R')2] [ 4a , R' = C≡C‐CH2‐NEt2; 4b , R' = C≡C‐t‐Bu]. In the solid state the mono‐adducts have clearly separated coordinatively saturated (coordination number four) and unsaturated aluminium atoms (coordination number three). In solution the groups R' show a fast exchange between both aluminium atoms as evident from the room temperature NMR spectra that showed in most cases equivalent CH(SiMe3)2 groups despite different coordination spheres of the metal atoms. Only 2b gave the expected splitting of resonances at ambient temperature, while cooling was required to prevent the dynamic process for 3a . The dialkynide 4a has a unique molecular structure with one of the lithium cations bonded to the α‐carbon atoms of the alkynido ligands and to the carbon atom of the methylene bridge which is five‐coordinate with a distorted trigonal bipyramidal coordination sphere.  相似文献   

13.
Molybdenum(VI) bis(imido) complexes [Mo(NtBu)2(LR)2] (R=H 1 a ; R=CF3 1 b ) combined with B(C6F5)3 ( 1 a /B(C6F5)3, 1 b /B(C6F5)3) exhibit a frustrated Lewis pair (FLP) character that can heterolytically split H−H, Si−H and O−H bonds. Cleavage of H2 and Et3SiH affords ion pairs [Mo(NtBu)(NHtBu)(LR)2][HB(C6F5)3] (R=H 2 a ; R=CF3 2 b ) composed of a Mo(VI) amido imido cation and a hydridoborate anion, while reaction with H2O leads to [Mo(NtBu)(NHtBu)(LR)2][(HO)B(C6F5)3] (R=H 3 a ; R=CF3 3 b ). Ion pairs 2 a and 2 b are catalysts for the hydrosilylation of aldehydes with triethylsilane, with 2 b being more active than 2 a . Mechanistic elucidation revealed insertion of the aldehyde into the B−H bond of [HB(C6F5)3]. We were able to isolate and fully characterize, including by single-crystal X-ray diffraction analysis, the inserted products Mo(NtBu)(NHtBu)(LR)2][{PhCH2O}B(C6F5)3] (R=H 4 a ; R=CF3 4 b ). Catalysis occurs at [HB(C6F5)3] while [Mo(NtBu)(NHtBu)(LR)2]+ (R=H or CF3) act as the cationic counterions. However, the striking difference in reactivity gives ample evidence that molybdenum cations behave as weakly coordinating cations (WCC).  相似文献   

14.
Recent advances in surface organometallic chemistry have enabled the detailed characterization of the surface species in single-site heterogeneous catalysts. However, the selective formation of bis-grafted surface species remains challenging because of the heterogeneity of the supporting surface. Herein, we introduce a metal complex bearing bidentate disilicate ligands, −OSi(OtBu)2OSi(OtBu)2O−, as a molecular precursor, which has a silicate framework adjacent to the metal (Pt) center. The grafting of the precursors on silica supports (MCM-41 and CARiACT Q10) proceeded through a substitution reaction on the silicon atoms of the disilicate ligand, which was verified by the detection of isobutene and tBuOH as the elimination products, to selectively yield bis-grafted surface species. The chemical structure of the surface species was characterized by solid-state NMR, and the chemical shift values of the ancillary ligands and 195Pt nuclei suggested that the bidentate coordination sphere was maintained following grafting.  相似文献   

15.
A series of zwitterionic aluminum complexes of the type AlX[(2‐O‐3,5‐tBu2C6H2)3PZ] (AlX [O3PZ]; X = Cl, Me, Et, and iBu; Z = H, Me) containing C3‐symmetric, formally dianionic, facially tridentate ligands [O3PZ]2? were prepared and structurally characterized. Although serendipitous, these complexes can be readily synthesized by partial protonolysis of AlX3 with equal molar (2‐HO‐3,5‐tBu2C6H2)3P (H3[O3P]) or [(2‐HO‐3,5‐tBu2C6H2)3p.m.e](OTf) ({H3[O3PMe]}OTf) in THF at 25°C or elevated temperatures. Alcoholysis of AlMe[O3PMe] ( 2 ) with an excess amount of MeOH in refluxing toluene generates AlOMe[O3PMe] ( 10 ). Salt metathesis of AlCl[O3PMe] ( 6 ) with nBuM (M = Li, MgCl) and NaOR (R = tBu, Ph) in ethereal solutions affords AlnBu[O3PMe] ( 9 ) and AlOR[O3PMe] (R = tBu ( 11 ), Ph ( 12 )), respectively. Reactivity of 10 , 11 , and 12 with respect to catalytic ring‐opening polymerization of ε‐caprolactone is assessed.  相似文献   

16.
While addition of [Cp2ReH] to [Bi(OtBu)3] leads to an equilibrium containing [Cp2Re‐Bi(OtBu)2], [{Cp2Re}2Bi(OtBu)], tBuOH and [CpRe(μη5,η1‐C5H4)Bi–ReCp2], in the presence of water [{(Cp2Re)2Bi}2O] ( 1 ) is formed selectively. Also [FpH] [Fp = (η5‐C5H5)(CO)2Fe] can be employed as a precursor to form heterometallic bismuth compounds. Synthesis of [FpBi{OCH(CF3)2}2]2 ( 5 ) can be achieved by reaction of [FpH] with [Bi{OCH(CF3)2}3(thf)]2 and carboxylates [FpBi(O2CR)2]2 are generated upon treatment of [FpH] with [Bi(O2CR)3] (R = CH3, tBu). While the compounds [Fp‐Bi(O2CR)2]2 can also be obtained from reactions with Fp‐Fp, they are formed far more readily using [FpH] as the precursor. They typically crystallize as dimers, like the alkoxide 5 . A monomeric compound of the type [Fp‐BiX2] ( 6 ) could be isolated for X = thd (tetramethylheptanedionate), that is, after the reaction of [FpH] with [Bi(thd)3]. Altogether, the results demonstrate the potential of [FpH] as a precursor for [Fp‐BiX2] compounds, which are formed in reactions with bismuth alkoxides, carboxylates and diketonates.  相似文献   

17.
The reactivity of the reduced anthracene complex of scandium [Li(thf)3][Sc{N(tBu)Xy}2(anth)] ( 2-anth-Li ) (Xy=3,5-Me2C6H3; anth=C14H102−, thf=tetrahydrofuran) toward Brønsted acid [NEt3H][BPh4] and azobenzene is reported. While a stepwise protonation of 2-anth-Li with two equivalents of [NEt3H][BPh4] afforded the scandium cation [Sc{N(tBu)Xy}2(thf)2][BPh4] ( 3 ), reduction of azobenzene gave a thermolabile, anionic scandium reduced azobenzene complex [Li(thf)][Sc{N(tBu)Xy}2(η2-PhNNPh)] ( 4 ), which slowly lost one of the anilide ligands to form the neutral scandium azobenzene complex dimer [Sc{N(tBu)Xy}(μ-η2:η2-Ph2N2)]2 ( 5 ). Exposure of 3 to CO2 produced the scandium carbamate complex [Sc{κ2-O2CN(tBu)(Xy)}2][BPh4] ( 6 ) as a result of CO2 insertion into the Sc−N bonds. In an attempt to prepare scandium hydrides, the reaction of 3 with the hydride sources LiAlH4 and Na[BEt3H] led to the terminal aluminum hydride [AlH{N(tBu)Xy}2(thf)] ( 7 ) and the scandium n-butoxide [Sc{N(tBu)(Xy)}2(μ-OnBu)] ( 8 ) after Sc/Al transmetalation and nucleophilic ring-opening of THF, respectively. All reported compounds isolated in moderate to good yields were fully characterized.  相似文献   

18.
The metathesis of [PhB(μ‐NtBu)2]AsCl and tBuN(H)Li in 1:1 molar ratio in diethyl ether produced the amido derivative [PhB(μ‐NtBu)2AsN(tBu)H] ( 1 ) in good yield. The lithiation of 1 with one equivalent of nBuLi afforded the lithium salt [PhB(μ‐NtBu)2AsN(tBu)Li] ( 2a ). Both 1 and 2a were characterized by multinuclear NMR spectroscopy. The crystal structure of 2a is comprised of a U‐shaped, centrosymmetric dimer in which the monomeric [PhB(μ‐NtBu)2AsN(tBu)]?Li+ units are linked by Li‐N interactions to give a six‐rung ladder. Oxidation of 2a with one‐half equivalent of I2 in diethyl ether resulted in hydrogen abstraction from the solvent to give the dimeric lithium iodide adduct {[PhB(μ‐NtBu)2AsN(tBu)H]LiI}2 ( 1 ·LiI) with a central Li2I2 ring.  相似文献   

19.
Enantiomerically pure triflones R1CH(R2)SO2CF3 have been synthesized starting from the corresponding chiral alcohols via thiols and trifluoromethylsulfanes. Key steps of the syntheses of the sulfanes are the photochemical trifluoromethylation of the thiols with CF3Hal (Hal=halide) or substitution of alkoxyphosphinediamines with CF3SSCF3. The deprotonation of RCH(Me)SO2CF3 (R=CH2Ph, iHex) with nBuLi with the formation of salts [RC(Me)? SO2CF3]Li and their electrophilic capture both occurred with high enantioselectivities. Displacement of the SO2CF3 group of (S)‐MeOCH2C(Me)(CH2Ph)SO2CF3 (95 % ee) by an ethyl group through the reaction with AlEt3 gave alkane MeOCH2C(Me)(CH2Ph)Et of 96 % ee. Racemization of salts [R1C(R2)SO2CF3]Li follows first‐order kinetics and is mainly an enthalpic process with small negative activation entropy as revealed by polarimetry and dynamic NMR (DNMR) spectroscopy. This is in accordance with a Cα? S bond rotation as the rate‐determining step. Lithium α‐(S)‐trifluoromethyl‐ and α‐(S)‐nonafluorobutylsulfonyl carbanion salts have a much higher racemization barrier than the corresponding α‐(S)‐tert‐butylsulfonyl carbanion salts. Whereas [PhCH2C(Me)SO2tBu]Li/DMPU (DMPU = dimethylpropylurea) has a half‐life of racemization at ?105 °C of 2.4 h, that of [PhCH2C(Me)SO2CF3]Li at ?78 °C is 30 d. DNMR spectroscopy of amides (PhCH2)2NSO2CF3 and (PhCH2)N(Ph)SO2CF3 gave N? S rotational barriers that seem to be distinctly higher than those of nonfluorinated sulfonamides. NMR spectroscopy of [PhCH2C(Ph)SO2R]M (M=Li, K, NBu4; R=CF3, tBu) shows for both salts a confinement of the negative charge mainly to the Cα atom and a significant benzylic stabilization that is weaker in the trifluoromethylsulfonyl carbanion. According to crystal structure analyses, the carbanions of salts {[PhCH2C(Ph)SO2CF3]Li? L }2 ( L =2 THF, tetramethylethylenediamine (TMEDA)) and [PhCH2C(Ph)SO2CF3]NBu4 have the typical chiral Cα? S conformation of α‐sulfonyl carbanions, planar Cα atoms, and short Cα? S bonds. Ab initio calculations of [MeC(Ph)SO2tBu]? and [MeC(Ph)SO2CF3]? showed for the fluorinated carbanion stronger nC→σ* and nO→σ* interactions and a weaker benzylic stabilization. According to natural bond orbital (NBO) calculations of [R1C(R2)SO2R]? (R=tBu, CF3) the nC→σ*S? R interaction is much stronger for R=CF3. Ab initio calculations gave for [MeC(Ph)SO2tBu]Li ? 2 Me2O an O,Li,Cα contact ion pair (CIP) and for [MeC(Ph)SO2CF3]Li ? 2 Me2O an O,Li,O CIP. According to cryoscopy, [PhCH2C(Ph)SO2CF3]Li, [iHexC(Me)SO2CF3]Li, and [PhCH2C(Ph)SO2CF3]NBu4 predominantly form monomers in tetrahydrofuran (THF) at ?108 °C. The NMR spectroscopic data of salts [R1(R2)SO2R3]Li (R3=tBu, CF3) indicate that the dominating monomeric CIPs are devoid of Cα? Li bonds.  相似文献   

20.
Treatment of [K(BIPMMesH)] (BIPMMes={C(PPh2NMes)2}2?; Mes=C6H2‐2,4,6‐Me3) with [UCl4(thf)3] (1 equiv) afforded [U(BIPMMesH)(Cl)3(thf)] ( 1 ), which generated [U(BIPMMes)(Cl)2(thf)2] ( 2 ), following treatment with benzyl potassium. Attempts to oxidise 2 resulted in intractable mixtures, ligand scrambling to give [U(BIPMMes)2] or the formation of [U(BIPMMesH)(O)2(Cl)(thf)] ( 3 ). The complex [U(BIPMDipp)(μ‐Cl)4(Li)2(OEt2)(tmeda)] ( 4 ) (BIPMDipp={C(PPh2NDipp)2}2?; Dipp=C6H3‐2,6‐iPr2; tmeda=N,N,N′,N′‐tetramethylethylenediamine) was prepared from [Li2(BIPMDipp)(tmeda)] and [UCl4(thf)3] and, following reflux in toluene, could be isolated as [U(BIPMDipp)(Cl)2(thf)2] ( 5 ). Treatment of 4 with iodine (0.5 equiv) afforded [U(BIPMDipp)(Cl)2(μ‐Cl)2(Li)(thf)2] ( 6 ). Complex 6 resists oxidation, and treating 4 or 5 with N‐oxides gives [{U(BIPMDippH)(O)2‐ (μ‐Cl)2Li(tmeda)] ( 7 ) and [{U(BIPMDippH)(O)2(μ‐Cl)}2] ( 8 ). Treatment of 4 with tBuOLi (3 equiv) and I2 (1 equiv) gives [U(BIPMDipp)(OtBu)3(I)] ( 9 ), which represents an exceptionally rare example of a crystallographically authenticated uranium(VI)–carbon σ bond. Although 9 appears sterically saturated, it decomposes over time to give [U(BIPMDipp)(OtBu)3]. Complex 4 reacts with PhCOtBu and Ph2CO to form [U(BIPMDipp)(μ‐Cl)4(Li)2(tmeda)(OCPhtBu)] ( 10 ) and [U(BIPMDipp)(Cl)(μ‐Cl)2(Li)(tmeda)(OCPh2)] ( 11 ). In contrast, complex 5 does not react with PhCOtBu and Ph2CO, which we attribute to steric blocking. However, complexes 5 and 6 react with PhCHO to afford (DippNPPh2)2C?C(H)Ph ( 12 ). Complex 9 does not react with PhCOtBu, Ph2CO or PhCHO; this is attributed to steric blocking. Theoretical calculations have enabled a qualitative bracketing of the extent of covalency in early‐metal carbenes as a function of metal, oxidation state and the number of phosphanyl substituents, revealing modest covalent contributions to U?C double bonds.  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