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 13 毫秒
1.
The deprotonation of the nido‐anion [B11H14] by two equivalents of LitBu yields the anion [B11H12]3–. Three observed 11B NMR shifts of this anion in the ratio 1 : 5 : 5 are in agreement with shifts calculated by the GIAO method on the basis of the ab initio computed geometry. The deprotonation can be reversed, giving back [B11H14] via [B11H13]2–. The thermolysis of [Li(thp)x]3[B11H12] in thp at 80 °C leads to the closo‐borate [Li(thp)3]2[B11H11] under elimination of LiH. Anhydrous air transforms [B11H12]3– into the known oxa‐nido‐dodecaborate [OB11H12]. The rhoda‐closo‐dodecaborate [L2RhB11H11]3– is formed from [B11H12]3– and RhL3Cl (L = PPh3).  相似文献   

2.
DFT‐calculations of the geometries of the closo‐anion [B11H11]2– in its ground state and in the transition state of its skeletal rearrangement and of the protonated species [B11H12] in its ground state were performed at the B3LYP/6‐31++G(d,p) level. The corresponding NMR shifts were computed on the basis of the optimized geometry by the GIAO method at the same level. Calculated and observed NMR data are in good agreement and thus prove the structure of [B11H12], previously deduced from 2 D‐NMR spectra. The addition of water, ethanol, and pyridine to [B11H12] at low temperature gave the nido‐species [B11H13(OH)], [B11H13(OEt)], and [B11H12(py)], respectively. The structures of these anions were investigated by NMR methods and the last two of them by crystal structure analyses of appropriate salts. The course of the addition reactions can be rationalized on the basis of the structurally characterized reaction components.  相似文献   

3.
The addition of neutral (L = py, NEt3, NHEt2, NH2tBu) and anionic Lewis bases (X = OH, Br, N3, Me, NHBu , NHtBu, [FeCp(CO)2]) to aza‐closo‐dodecaboranes RNB11H11 ( 1 ) or to derivatives thereof with boron bound non‐hydrogen ligands yields nido‐clusters RNB11H11L or [RNB11H11X] or derivatives thereof, respectively, the non‐planar pentagonal aperture N—B4—B9—B8—B5 of which hosts a B8—B9 hydrogen bridge. The base is either bound to B8 ( 3 )or B4 ( 5 )or B2( 7 ). The structures of these adducts are concluded from 1H and 11B NMR including 2D‐NMR spectra, and in the case of MeNB11H11(NHEt2) (type 3 ) also by a crystal structure analysis. With two of the adducts MeNB11H11L (L = py, NHEt2), isomers of the type 3 , 5 , and 7 , and with two of the adducts, MeNB11H11(NH2tBu) and {MeNB11H11[FeCp(CO)2]}, isomers of the type 3 and 7 could be identified. The position of boron‐bound ligands during the addition of bases to 1 shows, that only vertices of the ortho‐belt of 1 are involved in the opening process. A mechanism is made plausible that starts by the attack of the base at B2 of 1 and opening of the N‐B2 bond, denoted as a [3c, 1c]‐collocation, to give 2 with an endo‐H atom, whose migration into an adjacent bridge position and opening of a second B—N bond, denoted as a [3c, 2c]‐translocation, gives 3 ; this isomer can be transformed into 7 by a sequence of [3c, 2c]‐translocations via the isomers 4 , 5 , and 6 . The chiral type 3 species MeNB11H11L with L = NHEt2, NH2tBu undergo a rapid enantiomerization, for whose mechanism the exchange of the bridging and the amine‐H atom has been made plausible.  相似文献   

4.
Substitution of the dicarbaundecaborate anion nido‐7,8‐C2B9H12? ( 1 ) by precise hydride abstraction followed by nucleophilic attack usually leads to symmetric products 10‐R‐nido‐7,8‐C2B9H11. However, thioacetamide (MeC(S)NH2) as nucleophile and acetone/AlCl3 as hydride abstractor gave asymmetric 9‐[MeC(NHiPr)S]‐nido‐7,8‐C2B9H11 ( 2 ), whereas N,N‐dimethylthioacetamide (MeC(S)NMe2) gave the expected symmetric 10‐[MeC(NMe2)S]‐nido‐7,8‐C2B9H11 ( 4 ). For the formation of 2 , acetone and thioacetamide are assumed to give the intermediate MeC(S)N(CMe2) ( 3 ), which then attacks 1 with formation of 2 . Similarly, reaction of acetyliminium chloride [MeC(O)NH(CPh2)]Cl ( 5 ) with 1 in THF gave a mixture of 9‐ and 10‐substituted [MeC(NHCHPh2)O]‐nido‐7,8‐C2B9H11 ( 6 and 7 , respectively). These reactions are the first examples in which compounds (here heterodienes) that unite the functionalities of both hydride acceptor and nucleophilic site react with 1 in a bimolecular fashion. Furthermore, the analogous reaction of 1 and 5 (in an equilibrium mixture with acetyl chloride and benzophenone imine) in MeCN afforded 10‐[MeC(NCPh2)NH]‐nido‐7,8‐C2B9H11 ( 8 ) and MeC(O)NHCHPh2 ( 9 ).  相似文献   

5.
The closo‐dodecaborate [B12H12]2? is degraded at room temperature by oxygen in an acidic aqueous solution in the course of several weeks to give B(OH)3. The degradation is induced by Ag2+ ions, generated from Ag+ by the action of H2S2O8. Oxa‐nido‐dodecaborate(1?) is an intermediate anion, that can be separated from the reaction mixture as [NBzlEt3][OB11H12] after five days in a yield of 18 %. The action of FeCl3 on the closo‐undecaborate [B11H11]2? in an aqueous solution gives either [B22H22]2? (by fusion) or nido‐B11H13(OH)? (by protonation and hydration), depending on the concentration of FeCl3. In acetonitrile, however, [B11H11]2? is transformed into [OB11H12]? by Fe3+ and oxygen. The radical anions [B12H12] ˙ ? and [B11H11] ˙ ? are assumed to be the primary products of the oxidation with the one‐electron oxidants Ag2+ and Fe3+, respectively. These radical anions are subsequently transformed into [OB11H12]? by oxygen. The crystal structure analysis shows that the structure of [OB11H12]? is derived from the hypothetical closo‐oxaborane OB12H12 by removal of the B3 vertex, leaving a non‐planar pentagonal aperture with a three‐coordinate O vertex, as predicted by NMR spectra and theory.  相似文献   

6.
Tetraethyl­ammonium 7‐di­methyl­sulfanyl‐nido‐dodeca­hydro­undecaborate, [Et4N][7‐Me2S‐nido‐B11H12] or C8H20N+·C2H18B11S, is a product of the deprotonation of [7‐Me2S‐nido‐B11H13] with KHBEt3 and precipitation with tetraethyl­ammonium chloride. The effect of removing one endo‐terminal H atom is to cause a general contraction of the open‐face B—B distances.  相似文献   

7.
The potassium salt of the [1‐H2N‐2‐F‐closo‐1‐CB11H10] anion ( 1 ) was obtained from an insertion reaction of Li3[7‐H2N‐nido‐7‐CB10H10] with BF3 · OEt2. Anion 1 was protonated to the neutral species 1‐H3N‐2‐F‐closo‐1‐CB11H10 (H 1 ) and it was iodinated with ICl to the [1‐H2N‐2‐F‐closo‐1‐CB11I10] anion ( 2 ). All species were characterized by multinuclear NMR, IR, and Raman spectroscopy as well as by elemental analysis. The structure of H 1· (CH3)2CO was studied by single‐crystal X‐ray diffraction and the experimentally determined bond lengths are compared to values derived from density functional calculations.  相似文献   

8.
The Lanthanum Dodecahydro‐closo‐Dodecaborate Hydrate [La(H2O)9]2[B12H12]3·15 H2O and its Oxonium‐Chloride Derivative [La(H2O)9](H3O)Cl2[B12H12]·H2O By neutralization of an aqueous solution of the free acid (H3O)2[B12H12] with basic La2O3 and after isothermic evaporation colourless, face‐rich single crystals of a water‐rich lanthanum(III) dodecahydro‐closo‐dodecaborate hydrate [La(H2O)9]2[B12H12]3·15 H2O are isolated. The compound crystallizes in the trigonal system with the centrosymmetric space group (a = 1189.95(2), c = 7313.27(9) pm, c/a = 6.146; Z = 6; measuring temperature: 100 K). The crystal structure of [La(H2O)9]2[B12H12]3·15 H2O can be characterized by two of each other independent, one into another posed motives of lattice components. The [B12H12]2− anions (d(B–B) = 177–179 pm; d(B–H) = 105–116 pm) are arranged according to the samarium structure, while the La3+ cations are arranged according to the copper structure. The lanthanum cations are coordinated in first sphere by nine oxygen atoms from water molecules in form of a threecapped trigonal prism (d(La–O) = 251–262 pm). A coordinative influence of the [B12H12]2− anions on La3+ has not been determined. Since “zeolitic” water of hydratation is also present, obviously the classical H–Oδ–···H–O‐hydrogen bonds play a significant role in the stabilization of the crystal structure. During the conversion of an aqueous solution of (H3O)2[B12H12] with lanthanum trichloride an anion‐mixed salt with the composition [La(H2O)9](H3O)Cl2[B12H12]·H2O is obtained. The compound crystallizes in the hexagonal system with the non‐centrosymmetric space group (a = 808.84(3), c = 2064.51(8) pm, c/a = 2.552; Z = 2; measuring temperature: 293 K). The crystal structure can be characterized as a layer‐like structure, in which [B12H12]2− anions and H3O+ cations alternate with layers of [La(H2O)9]3+ cations (d(La–O) = 252–260 pm) and Cl anions along [001]. The [B12H12]2− (d(B–B) = 176–179 pm; d(B–H) = 104–113 pm) and Cl anions exhibit no coordinative influence on La3+. Hydrogen bonds are formed between the H3O+ cations and [B12H12]2− anions, also between the water molecules of [La(H2O)9]3+ and Cl anions, which contribute to the stabilization of the crystal structure.  相似文献   

9.
The closo‐undecaborate A2[B11H11] (A = NBzlEt3) can be halogenated with excess N‐chlorosuccine imide, bromine or iodine, respectively, to give the perhalo‐closo‐undecaborates A2[B11Hal11] (Hal = Cl, Br, I). The chlorination in the 11 : 1 ratio of the reagents yields A2[B11HCl10], whose subsequent iodination makes A2[B11Cl10I] available. The three type [B11Hal11]2– anions show only one and the two type [B11Cl10X]2– anions (X = H, I) only two 11B NMR peaks in the ratio 10 : 1, thus exhibiting the same degenerate rearrangement of the octadecahedral B11 skeleton as is well‐known for [B11H11]2–. The crystal structure analysis of A2[B11Br11] and A2[B11I11] reveals a rigid octadecahedral skeleton in the solid state, up to 330 K, whose B–B bond lengths deviate more or less from the idealized C2v gas phase structure, but are in good accordance with the distances of A2[B11H11]. Electrochemical experiments elucidate the mechanism of the known oxidation of [B11H11]2– to give [B22H22]2–: A first one‐electron transfer is followed by the dimerization of the [B11H11] monoanion, whereas neutral B11H11, a presumably most reactive species, does not play a role as an intermediate. The electrochemical oxidation of [B11Hal11]2– anions also starts with a one‐electron transfer, which is perfectly reversible only in the case of Hal = Br. There is no electrochemical indication for the formation of [B22Hal22]2–. The neutral species B11Hal11 should be a short‐lived, very reactive species.  相似文献   

10.
The 6‐aza‐nido‐decaboranes RNB9H11 ( 1a—d ; R = H, Ph, 4‐C6H4Me, 4‐C6H4Cl) act as 1, 2‐hydroboration agents via their 9‐BH vertex, giving products RNB9H10R′. The boranes 1a, b and 3‐hexyne yield the 9‐(1‐ethyl‐1‐butenyl)‐6‐aza‐nido‐decaboranes 2a, b (R′ = CEt = CHEt). 2, 3‐Dimethyl‐2‐butene is hydroborated by 1a—d under formation of the 9‐(1, 1, 2‐trimethylpropyl)‐6‐aza‐nido‐decaboranes 3a—d (R′ = —CMe2 —CHMe2). With the boranes 1a—c and (trimethylsilyl)ethene, a 85:15 mixture of the products (RNB9H10)CH2CH2(SiMe3)( 4a—c ) and their chiral isomers (RNB9H10)CH(SiMe3)CH3 ( 5a—c ) is obtained. The action of BH3(SMe2) on the mixtures 4b/5b or 4c/5c results in a closure of the nido‐NB9 skeleton of 4b or 4c , respectively, with a closo‐NB11 skeleton of the products RNB11H10R′ ( 6b or 6c;R′ = CH2CH2(SiMe3)); R′ is found in position 7 of 6b, c . All products of the type 2—6 are characterised by NMR.  相似文献   

11.
The OCO carboxylate unit of pivalic acid adds to the B–B bond of the azadiboriridine NB2R3 ( 1 a , R = tBu) to give the chiral heterocyclohexadiene 2 a ; the enantiomers of 2 a are transformed into one another by a [1,3] sigmatropic hydride transfer along the B–N–B ring fragment. The azadiboracyclopentanes 3 a – e are formed from 1 a and the alkenes ethene, propene, isobutene, (trimethylsilyl)ethene, and 2,3‐dimethyl‐1‐butene. Only one double bond of cyclopentadiene and 1,3‐butadiene reacts in the same way to give 3 f , g , respectively, and both of the double bonds of 1,3‐butadiene react with an excess of 1 a to give 3 h , which is obtained in a 9 : 1 mixture of racemate and meso‐isomer; the meso‐isomer crystallizes in the space group P21/n. The corresponding diazadiboracyclopentane 3 i and the triazadiboracyclopentane 3 j are formed from 1 a and N‐phenyl benzaldimine or azobenzene, respectively. Ethyne and 1 a give either the azadiboracyclopentene 4 a (1 : 1) or the diazatetraborabicyclo[3.3.0]octane 3 k (1 : 2). The phosphaalkyne P≡C–tBu and 1 a  analogously yield the heterocyclopentene 4 c . The insertion of SitBu2 into 1 a to give the azasiladiboracyclobutane 5 a is achieved by applying Li powder and tBu2SiCl2. The hitherto unknown azadiboriridines BN2R2R′ (R = tBu; R′ = 1‐iPr, 2‐Mes, 2‐CMe2Et: 1 b – d ) were synthesized by the chloroboration of the iminoboranes RB≡NiPr and RB≡NR with RBCl2, MesBCl2, and (EtMe2C)BCl2, respectively, and subsequent dechlorination of the isolated and characterized diborylamines Cl–BR–NiPr–BR–Cl ( 6 a ), Cl–BR–NR–BMes–Cl ( 6 b ), and Cl–BR–NR–B(CMe2Et)–Cl ( 6 c ), respectively, with lithium (Mes = mesityl).The azadiboriridine 1 b dimerizes to give the diaza‐nido‐hexaborane 7 a , whereas 1 c and 1 d are storable at room temperature. The product 1 c crystallizes as a racemate in the space group P21/c; its ring geometry differs from that of the known N‐mesityl isomer.  相似文献   

12.
The reaction of [PtCl2(PPh3)2] with closo‐B10H102? in ethanol under reflux conditions gave two nido 11‐vertex platinaundecaborane clusters: [(PPh3)2PtB10H10‐8,10‐(OEt)2]·CH2Cl2 (1) and [(PPh3)2PtB10H11‐11‐OEt]·CH2Cl2 (2) . A novel B10H102? deboronated nido 11‐vertex diplatinaundecaborane [(µ‐PPh2)(PPh3)2Pt2B9H6‐3,9,11‐(OEt)3]·CH2Cl2 (3) was obtained when the same reaction was carried out under solvothermal conditions. All of these compounds were characterized by infrared spectroscopy, NMR spectroscopy, elemental analysis and single‐crystal X‐ray diffraction. Both clusters 1 and 2 have a nido 11‐vertex {PtB10} polyhedral skeleton in which the Pt atom lies in the open PtB4 face. Each Pt atom connects with four B atoms and two P atoms of the PPh3 ligands. Cluster 3 has a nido 11‐vertex {Pt2B9} polyhedral skeleton in which two Pt atoms sit in neighbouring positions of the open Pt2B3 face, bridged by a PPh2 group. Each Pt atom connects three B atoms and a P atom of the PPh3 ligand. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

13.
ChemInform is a weekly Abstracting Service, delivering concise information at a glance that was extracted from about 100 leading journals. To access a ChemInform Abstract of an article which was published elsewhere, please select a “Full Text” option. The original article is trackable via the “References” option.  相似文献   

14.
Bis(tetramethylammonium) dodecahydrododecaborate, [(CH3)4N]2[B12H12], and bis(tetramethylammonium) dodecahydrododecaborate acetonitrile, [(CH3)4N]2[B12H12] · CH3CN, were synthesized and characterized via Infrared, 1H and 11B NMR spectroscopy. [(CH3)4N]2[B12H12] crystallizes isopunctual to the alkali metal dodecaborates. The crystal structure of [(CH3)4N]2[B12H12] · CH3CN was determined from single crystal data and refined in the orthorhombic crystal system (Pcmn, no. 62, a = 898.68(8), b = 1312.85(9) c = 1994.5(1) pm, R(|F| , 4σ) = 5.9%, wR(F2) = 18.3%). Here, the geometry of the dodecaborate anion is that of an almost ideal icosahedron, less distorted than most other dodecaborates known. By low‐temperature Guinier‐Simon diffractometry phase transitions were detected for [(CH3)4N]2[B12H12] and [(CH3)4N]2[B12H12] · CH3CN at –70 and –15 °C, respectively.  相似文献   

15.
The tetranion ligand η7-C2B10H124− has been observed for the first time in 1 , which was obtained from the reaction of o-C2B10H12 with excess K metal followed by treatment with UCl4. As shown in the picture (without K cations and coordinated THF molecules), 1 is a centrosymmetric dimer with a bent sandwich structure.  相似文献   

16.
L,L ‐lactide (LA) and ε‐caprolactone (CL) block copolymers have been prepared by initiating the poly(ε‐caprolactone) (PCL) block growth with living poly(L,L ‐lactide) (PLA*). In the previous attempts to prepare block copolymers this way only random copolyesters were obtained because the PLA* + CL cross‐propagation rate was lower than that of the PLA–CL* + PLA transesterification. The present paper shows that application of Al‐alkoxide active centers that bear bulky diphenolate ligands results in efficient suppression of the transesterification. Thus, the corresponding well‐defined di‐ and triblock copolymers could be prepared.

  相似文献   


17.
Transparent and nearly colorless single crystals of r‐LiB13C2 were obtained by reaction of boron with Li2CO3 in a Cu melt at 1250–1300 °C. The structure analysis [R3 m, a = 5.6535(1), c = 12.5320(2) Å, 421 independent reflections, 22 parameters, R1 = 0.034, wR2 = 0.093] revealed a crystal structure that can be described as a filling variant of rhombohedral B13C2. Li+ is located in a void above or below the linear CBC unit. The site occupation is close to 50 % resulting in an electron‐precise composition according to Wade's rules if a positive charge is given to the CBC entity: Li+(B12)2–(CBC)+. The displacement parameters of the CBC unit indicate disorder in the [001] direction, that relates to the short Li–C distance and the partial occupation of the Li+ site. The composition is confirmed by EELS measurements of single crystals. Band gap calculations give a value of 2.94 eV, which is in agreement to the crystals being colorless. The evaluation of the electron density by application of the QTAIM formalism as proposed by Bader modifies the assignment pictured above according to Wade's rules. In agreement to the electronegativities the carbon atoms carry a negative charge (–2.31/–2.42) and the effective charges are: Li+0.81(B12)+2.02(CBC)–2.83.  相似文献   

18.
Total syntheses of iso‐cladospolide B ( 1 ) and the 12‐membered macrolactone (6S,12R)‐6‐hydroxy‐12‐methyloxacyclododecane‐2,5‐dione ( 2 ), a non‐natural product, were achieved from a common intermediate starting from commercially available 1,9‐nonane diol.  相似文献   

19.
An isomer of B20H26, isolated from the autolysis of nido‐B10H14 in a silent electrical discharge, is shown to be the title compound 1,1′‐bis(nido‐decaboranyl). The mol­ecule has crystallographic inversion symmetry and a long intercage B—B bond of 1.704 (3) Å.  相似文献   

20.
Synthesis of several 4‐benzhydrylidenepiperidine analogs has been established starting from different (4‐phenylpiperidin‐4‐yl)‐arylmethanols via boron trifluoride etherate mediated rearrangement. The possible rearranged mechanism was proposed. Boron trifluoride etherate‐mediated rearrangement of the related derivatives was also examined. It presents a novel rearrangement reaction catalyzed by boron trifluoride etherate and broadens the scope of application.  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