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.
Halogenation of nido-B10H14 with C2H2Cl4, C2Cl6, Br2, or I2, produces by cluster degradation the (2 n)-closo-clusters B9X9 (X = Cl, Br, I). The synthesis of salts of the perhalogenated radical anions of the type (2 n + 1)-closo-[B9X9]· – and of the corresponding dianions (2 n + 2)-closo-[B9X9]2– from neutral B9X9 is described [n is the number of cluster atoms; (2 n), (2 n + 1), and (2 n + 2) is the number of cluster electrons]. Molecular and crystal structures of B9Cl9, B9Br9, [(C6H5)4P][B9Br9] · CH2Cl2, and [(C4H9)4N]2[B9Br9] · CH2Cl2 have been determined via X-ray diffraction. All three oxidation states of the cluster retain the tricapped trigonal prism. The reduction of the clusters B9X9 was shown by cyclic voltammetry in CH2Cl2 to proceed via two successive one-electron reversible steps, separated by at least 0.4 V. The paramagnetic radical anions [B9X9]· – (X = Cl, Br) were further characterized by magnetic susceptibility measurements of [Cp2Fe][B9X9] and [Cp2Co][B9X9], respectively. The EPR spectra of [B9X9]· – (X = Cl, Br, I) in glassy frozen CH2Cl2 solutions showed increasing g anisotropy for the heavier halogen derivatives, illustrating significant halogen participation at the singly occupied MO. The 11B NMR spectra of CD2Cl2 solutions of the neutral clusters B9X9 exhibit only one sharp resonance, indicating that the boron atoms are highly fluxional in solution. In contrast, two different boron resonances as expected for a rigid tricapped trigonal prism are clearly observed for the [B9X9]2– dianions in solutions and for solid B9Br9 in the 11B MAS NMR spectra. Temperature dependent 11B MAS NMR experiments on B9Br9 and [B9Br9]2– in the solid state show a reversible coalescence of the two resonances at higher temperature. 11B MAS NMR spectra and DTA measurements of [B9Br9]2– showed a phase transition.  相似文献   

2.
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   

3.
The Crystal Structures of the Dicesium Dodecahalogeno-closo-Dodecaborates Cs2[B12X12] (X = Cl, Br, I) and their Hydrates The perhalogenated derivatives Cs2[B12X12] (X = Cl - I) have been synthesized by reaction of Cs2[B12H12] with the respective elemental halogens (Cl2, Br2 and I2). Upon recrystallization from aqueous solution colourless, face-rich single crystals of the dihydrates (Cs2[B12X12] · 2 H2O) are obtained first which can be dehydrated topotactically via the monohydrates (Cs2[B12X12] · H2O) leaving to the solvent-free compounds (Cs2[B12X12]) behind without loss of their crystallinity. The ionic cesium salts were characterized by single crystal X-ray diffraction. All three halogenoborates are isostructural and they crystallize at room temperature in the trigonal space group (Cs2[B12Cl12]: a = 959.67(3) pm, c = 4564.2(2) pm; Cs2[B12Br12]: a = 997.92(3) pm, c = 4766.4(3) pm; Cs2[B12I12]: a = 1047.05(4) pm, c = 5018.3(3) pm; Z = 6). The crystal structures consist of a cubic closest packed host lattice formed by two crystallographically inequivalent quasi-icosahedral [B12X12]2- anions (Cs2[B12Cl12]: d(B-B) = 178 - 179 pm, d(B-Cl) = 179 - 180 pm; Cs2[B12Br12]: d(B-B) = 176 - 180 pm, d(B-Br) = 195 - 197 pm; Cs2[B12I12]: d(B-B) = 177 - 182 pm, d(B-I) = 214 - 217 pm). By ordered occupation of half of the tetrahedral and formally all octahedral interstices in every intermediate layer with Cs+ cations, a structure emerges where (Cs1)+ is trigonally non-planar coordinated by three (CN = 9) and (Cs2)+ tetrahedrally coordinated by four (CN = 12) [B12X12]2- anions. Thereby triangular faces of halogen atoms of the icosahedral clusters are coordinatively effective in both cases. In their mono- and dihydrates the incomplete coordination sphere of (Cs1)+ is completed by one and two water molecules, respectively. The thermal decomposition of the dicesium dodecahalogeno-closo-dodecaborate hydrates and their dehydration products was investigated using DTA/TG methods in a temperature range between room temperature and 1200 °C. Additionally the compounds were also characterized by 11B-NMR spectroscopy in aqueous solution.  相似文献   

4.
Chemical and Cyclovoltammetric Investigation of the Redoxreactions of the Decahalodecaborates closo ‐[B10X10]2– and hypercloso ‐[B10X10]· – (X = Cl, Br)1). Crystal Structure Analysis of Cs2[B10Br10] · 2 H2O The oxidation of the decachloro‐closo‐decaborates(2–) Cs2[B10Cl10] or [Me4N]2[B10Cl10] with Tl(CF3COO)3 leads to the corresponding radical monoanion hypercloso‐[B10Cl10] · –, which was characterized by ESR and UV/Vis spectroscopy. [B10Cl10] · – does not dimerize like [B10H10] · – but it is reduced by acetonitrile to the dianion [B10Cl10]2–. Cs2[B10Cl10] reacts with stronger oxidation agents like CoF3 (in dichloromethane) or XeF2 (in perfluorhexane), respectively, to yield B9Cl9 and, in traces, B8Cl8. In opposite to this, the decabromoderivative Cs2[B10Br10] does not show any reaction with Tl(CF3COO)3 in acetonitrile or with CoF3 in CH2Cl2. The oxidation of the dianions [B10X10]2– (X = Cl, Br) was studied by electroanalytical methods (cyclic voltammetry, chronoamperometry, chronocoulometry). Formal potentials were determined for the two steps of the reaction, which do not seem to be affected by structural rearrangements. The crystal structure of Cs2[B10Br10] · 2 H2O was analyzed by single‐crystal X‐ray diffraction. Cs2[B10Br10] · 2 H2O crystallizes monoclinic (space group I2/a, (no. 15), Z = 8, a = 1361.54(9) pm, b = 1215.89(5) pm, c = 3108.4(2) pm, α = 90°, β = 97.916(8)°, γ = 90°). The closo‐cluster B10Br102– has a bicapped square antiprismatic structure with idealized D4d symmetry.  相似文献   

5.
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   

6.
Electrophilic anions of type [B12X11] posses a vacant positive boron binding site within the anion. In a comparatitve experimental and theoretical study, the reactivity of [B12X11] with X=F, Cl, Br, I, CN is characterized towards different nucleophiles: (i) noble gases (NGs) as σ-donors and (ii) CO/N2 as σ-donor-π-acceptors. Temperature-dependent formation of [B12X11NG] indicates the enthalpy order (X=CN)>(X=Cl)≈(X=Br)>(X=I)≈(X=F) almost independent of the NG in good agreement with calculated trends. The observed order is explained by an interplay of the electron deficiency of the vacant boron site in [B12X11] and steric effects. The binding of CO and N2 to [B12X11] is significantly stronger. The B3LYP 0 K attachment enthapies follow the order (X=F)>(X=CN)>(X=Cl)>(X=Br)>(X=I), in contrast to the NG series. The bonding motifs of [B12X11CO] and [B12X11N2] were characterized using cryogenic ion trap vibrational spectroscopy by focusing on the CO and N2 stretching frequencies and , respectively. Observed shifts of and are explained by an interplay between electrostatic effects (blue shift), due to the positive partial charge, and by π-backdonation (red shift). Energy decomposition analysis and analysis of natural orbitals for chemical valence support all conclusions based on the experimental results. This establishes a rational understanding of [B12X11] reactivety dependent on the substituent X and provides first systematic data on π-backdonation from delocalized σ-electron systems of closo-borate anions.  相似文献   

7.
The reaction of germa‐closo‐dodecaborate with oneequivalent of silver halide AgX (X = Cl, Br) leads to the tetrameric1:1 adducts [Et3MeN]8[{AgCl(GeB11H11)}4] ( 1 ) and [Et3MeN]8[{AgBr(GeB11H11)}4] ( 2 ). A cubane‐like structure was determined in the solid state by single‐crystal X‐ray diffraction. The compounds were characterized by crystal structure analysis, 11B NMR spectroscopy and elemental analysis.  相似文献   

8.
Synthesis and Crystal Structure of Cadmium Dodecahydro closo‐Dodecaborate Hexahydrate, Cd(H2O)6[B12H12] Through neutralization of the aqueous free acid (H3O)2[B12H12] with cadmium carbonate (CdCO3) and after isothermic evaporation of the resulting solution, colourless lath‐shaped single crystals of Cd(H2O)6[B12H12] are obtained. Cadmium dodecahydro closo‐dodecaborate hexahydrate crystallizes at room temperature in the monoclinic system (space group: C2/m) with the lattice constants a = 1413.42(9), b = 1439.57(9), c = 749.21(5) pm and β = 97.232(4)° (Z = 4). The crystal structure of Cd(H2O)6[B12H12] can be regarded as a monoclinic distortion variant of the CsCl‐type structure. Two crystallographically different [Cd(H2O)6]2+ octahedra (d(Cd–O) = 227–230 pm) are present which only differ in their relative orientation. The intramolecular bond lengths for the quasi‐icosahedral [B12H12]2? cluster anions range in the intervals usually found for dodecahydro closo‐dodecaborates (d(B–B) = 177–179 pm, d(B–H) = 103–116 pm). The hydrogen atoms of the [B12H12]2? clusters have no direct coordinative influence on the Cd2+ cations. Due to the fact that no “zeolitic” crystal water molecules are present, a stabilization of the lattice takes place mainly via the B–Hδ?···H–O hydrogen bonds.  相似文献   

9.
Preparation and Crystal Structures of Dipyridiniomethane Monohalogenohydro-closo-Dodecaborates(2?), [(C5H5N)2CH2][B12H11X]; X = Cl, Br, I [B12H12]2? reacts with dihalogenomethanes CH2X2 in presence of trifluoro acetic acid, yielding the monohalogenododecaborates [B12H11X]2? (X = Cl, Br, I), which are separated by ion exchange chromatography on diethylaminoethyl(DEAE) cellulose from the starting compound and higher halogenated products. The X-ray structure determinations of [(C5H5N)2CH2][B12H11Cl] · 2(CH3)2SO (orthorhombic, space group Pnma, a = 17.351(6), b = 16.034(5), c = 9.659(2) Å, Z = 4) and of the isotypic bromo and iodo compounds [(C5H5N)2CH2][B12H11X] (monoclinic, space group P21/n, Z = 4; for X = Br: a = 7.339(2), b = 15.275(3), c = 16.761(4) Å, β = 96.80(2)°; for X = I: a = 7.4436(8), b = 15.3510(8), c = 16.9213(16) Å, ß = 97.326(7)°) exhibit crystal lattices build up by columns of substituted boron clusters and angular dications [(C5H5N)2CH2]2+ orientated along the shortest axis which are assembled to alternating layers.  相似文献   

10.
Chemical reduction of B9X9 (X = Cl, Br, I) with gaseous HI proceeds stepwise to give the neutral paramagnetic clusters HB9X9 · , and the corresponding diamagnetic clusters H2B9X9. Together they comprise the first neutral derivatives in the series BnHn+1 and BnHn+2 with n = 9. The EPR spectra of the paramagnetic HB9X9 · (X = Cl, Br, I) in glassy frozen CH2Cl2 solutions showed increasing g anisotropy for the heavier halogen derivatives, illustrating significant halogen participation at the singly occupied MO due to the larger spin-orbit coupling contributions. Temperature dependent 1H NMR spectra of H2B9X9 (in CD3CN, X = Cl, Br) indicate the presence of H2B9X9, [HB9X9], and [CD3CNH]+ with H2B9X9 acting as a Brønsted acid. The corresponding 11B NMR spectra (in CD3CN) show the presence of the dianions [B9X9]2– as a result of the protonation of CD3CN. The 11B resonances of the species H2B9X9 and [HB9X9] are obscured by superimposition of the two resonance lines of the dianions [B9X9]2–. Temperature dependent 11B{1H} MAS-NMR spectra of H2B9Br9 show coalescence at 410 K and hence dynamic behaviour of the neutral B9-cluster in the solid. Cyclic voltammetry experiments of H2B9Br9 in CH3CN solvent) are compatible with the redox sequence [B9Br9]2––[B9Br9] · ––B9Br9. Quantum chemical calculations with the electron localization function (ELF) are described.  相似文献   

11.
Preparation and Vibrational Spectra of trans-[Pt(acac)2X2] (X ? Cl, Br, I, SCN, SeCN, N3) By electrolytical oxidation of [Pt(acac)2] in presence of chloride or bromide, dissolved in dichlormethane, trans-[Pt(acac)2X2], X ? Cl, Br, are formed. On treatment of trans-[Pt(acac)2I2] with silver pseudohalides trans-[Pt(acac)2X2], X ? SCN, SeCN, N3, are obtained. Beside the nearly persistent bands of coordinated acetylacetonate in the Raman spectra the intensive and sharp symmetric, in the IR spectra the corresponding antisymmetric stretching vibration of the X? Pt? X axis is observed. The observance of the rule of mutual exclusion proves the complexes to belong to point group D2h. From the resonance Raman spectrum of trans-[Pt(acac)2I2] for vs (Pt? I), Ag, the harmonic frequency ω1 = 142.45 cm?1 and the inharmonicity constant x11 = 0.48 cm?1 is calculated. In the Raman spectrum of trans-[Pt(acac)2Cl2] vs (Pt? Cl) is splitted by the isotops 35Cl/37Cl into the triplet 340, 335, 330 cm?1 giving the force constant fPtCl = 2.01 N/cm.  相似文献   

12.
Transformation of [W6X8]X4 + 3 X2 = [W6X12]X6 (X = Cl, Br) The transformation of [W6X8]X4 + 3 X2 = [W6X12]X6 (X = Cl, Br) has been investigated by changing the relation Cl2/Br2 and the temperature. In this way the compounds [W6Br12?nCln]Cl6?mBrm are isolated. All of the products are isotypic with W6Cl18 and W6Br18. Most often n equals 6, however compounds with other relations of Cl/Br are also observed (e. g. n = 4.8) The 6 ligands standing outside of the brackets are replaced by Cl or Br. The substitution of [W6Br6Cl6]Cl6 by means of bromine leads to the cluster [W6Br12]X6. The backward transformation of the cluster compound [W6Br12]Br6 happens by decomposition on the thermobalance, e. g. according to Gl. (1) (See Inhaltsübersicht). By analogy [W6Br12]Cl6 is decomposed to [W6Br8]Cl2Br2, which by treatment with conc. HCl is transformed into [W6Br8]Cl4 · 2 H2O.  相似文献   

13.
Investigations on the Crystal Structure of Lithium Dodecahydro‐closo‐dodecaborate from Aqueous Solution: Li2(H2O)7[B12H12] By neutralization of an aqueous solution of the acid (H3O)2[B12H12] with lithium hydroxide (LiOH) and subsequent isothermic evaporation of the resulting solution to dryness, it was possible to obtain the heptahydrate of lithium dodecahydro‐closo‐dodecaborate Li2[B12H12] · 7 H2O (≡ Li2(H2O)7[B12H12]). Its structure has been determined from X‐ray single crystal data at room temperature. The compound crystallizes as colourless, lath‐shaped, deliquescent crystals in the orthorhombic space group Cmcm with the lattice constants a = 1215.18(7), b = 934.31(5), c = 1444.03(9) pm and four formula units in the unit cell. The crystal structure of Li2(H2O)7[B12H12] can not be described as a simple AB2‐structure type. Instead it forms a layer‐like structure analogous to the well‐known barium compound Ba(H2O)6[B12H12]. Characteristic feature is the formation of isolated cation pairs [Li2(H2O)7]2+ in which the water molecules form two [Li(H2O)4]+ tetrahedra with eclipsed conformation, linked to a dimer via a common corner. The bridging oxygen atom (∢(Li‐ O ‐Li) = 112°) thereby formally substitutes Ba2+ in Ba(H2O)6[B12H12] according to (H2 O )Li2(H2O)6[B12H12]. A direct coordinative influence of the [B12H12]2— cluster anions to the Li+ cations is not noticeable, however. The positions of the hydrogen atoms of both the water molecules and the [B12H12]2— units have all been localized. In addition, the formation of B‐Hδ—···δ+H‐O‐hydrogen bonds between the water molecules and the hydrogen atoms from the anionic [B12H12]2— clusters is considered and their range and strength is discussed. The dehydratation of the heptahydrate has been investigated by DTA‐TG measurements and shown to take place in two steps at 56 and 151 °C, respectively. Thermal treatment leads to the anhydrous lithium dodecahydro‐closo‐dodecaborate Li2[B12H12], eventually.  相似文献   

14.
The Electron Localization Function in closo Boron Clusters The structure and the electron density in the closo boron clusters B4X4 (X = H, Cl, Br, I), B6X62? (X = H, Cl, Br, I) and B12H122? were determined by pseudopotential Hartree-Fock calculations. The Electron Localization Function (ELF) was used to interpret the bonding characteristics. The regions of high ELF values in all cases have the form of the dual polyhedron of the boron cage. They show perfectly the 3 center 2 electron bonds. The comparison between Hartree-Fock and Extended Hückel calculations point out that semiempirical calculations can also be a good basis for ELF interpretations.  相似文献   

15.
On the Crystal Structures of the Transition‐Metal(II) Dodecahydro‐closo‐Dodecaborate Hydrates Cu(H2O)5.5[B12H12]·2.5 H2O and Zn(H2O)6[B12H12]·6 H2O By neutralization of an aqueous solution of the free acid (H3O)2[B12H12] with basic copper(II) carbonate or zinc carbonate, blue lath‐shaped single crystals of the octahydrate Cu[B12H12]·8 H2O (≡ Cu(H2O)5.5[B12H12]·2.5 H2O) and colourless face‐rich single crystals of the dodecahydrate Zn[B12H12]·12 H2O (≡ Zn(H2O)6[B12H12]·6 H2O) could be isolated after isothermic evaporation. Copper(II) dodecahydro‐closo‐dodecaborate octahydrate crystallizes at room temperature in the monoclinic system with the non‐centrosymmetric space group Pm (Cu(H2O)5.5[B12H12]·2.5 H2O: a = 768.23(5), b = 1434.48(9), c = 777.31(5) pm, β = 90.894(6)°; Z = 2), whereas zinc dodecahydro‐closo‐dodecaborate dodecahydrate crystallizes cubic in the likewise non‐centrosymmetric space group F23 (Zn(H2O)6[B12H12]·6 H2O: a = 1637.43(9) pm; Z = 8). The crystal structure of Cu(H2O)5.5[B12H12]·2.5 H2O can be described as a monoclinic distortion variant of the CsCl‐type arrangement. As characteristic feature the formation of isolated [Cu2(H2O)11]4+ units as a condensate of two corner‐linked Jahn‐Teller distorted [Cu(H2O)6]2+ octahedra via an oxygen atom of crystal water can be considered. Since “zeolitic” water of hydratation is also present, obviously both classical H–Oδ?···H–O and non‐classical B–Hδ?···H–O hydrogen bonds play a significant role for the stabilization of the structure. A direct coordinative influence of the quasi‐icosahedral [B12H12]2? anions on the Cu2+ cations has not been determined. The zinc compound Zn(H2O)6[B12H12]·6 H2O crystallizes in a NaTl‐type related structure. Two crystallographically different [Zn(H2O)6]2+ octahedra are present, which only differ in their relative orientation within the packing of the [B12H12]2? anions. The stabilization of the crystal structure takes place mainly via H–Oδ?···H–O hydrogen bonds, since again the hydrogen atoms of the [B12H12]2? anions have no direct coordinative influence on the Zn2+ cations.  相似文献   

16.
Nitro-functionalized undecahalogenated closo-dodecaborates [B12X11(NO2)]2− were synthesized in high purities and characterized by NMR, IR, and Raman spectroscopy, single crystal X-diffraction, mass spectrometry, and gas-phase ion vibrational spectroscopy. The NO2 substituent leads to an enhanced electronic and electrochemical stability compared to the parent perhalogenated [B12X12]2− (X=F–I) dianions evidenced by photoelectron spectroscopy, cyclic voltammetry, and quantum-chemical calculations. The stabilizing effect decreases from X=F to X=I. Thermogravimetric measurements of the salts indicate the loss of the nitric oxide radical (NO.). The homolytic NO. elimination from the dianion under very soft collisional excitation in gas-phase ion experiments results in the formation of the radical [B12X11O]2−.. Theoretical investigations suggest that the loss of NO. proceeds via the rearrangement product [B12X11(ONO)]2−. The O-bonded nitrosooxy structure is thermodynamically more stable than the N-bonded nitro structure and its formation by radical recombination of [B12X11O]2−. and NO. is demonstrated.  相似文献   

17.
Preparation of trans-[Pt(ox)2X2]2? (X = Cl, Br, I, SCN, OH) by Oxidative Addition to [Pt(ox)2]2? in Organic Solvents After extraction of [Pt(ox)2]2? with long-chain alkyl-ammonium ions into organic solvents the new PtIV complexes trans-[Pt(ox)2X2]2?, X = Cl, Br, I, SCN, OH, are formed directly by oxidative addition. In nonpolar solvents the bulky organic cations prevent the formation of compounds with columnar structure which by partial oxidation in aqueous solution are formed immediately. The IR and Ra spectra of the stable anhydrous (TBA) salts are assigned according to point group D2h. A characteristical dependence of the C?O, C? O, and Pt? O stretching modes in response to the oxidation state of the central ion is observed. There is vibrational fine structure in the absorption spectrum of [Pt(ox)2]2? measured at 10 K with long progressions by coupling of d—d transitions with vs(Pt? O) and vs(C?O). The characteristical feature in the UV/VIS spectra of the PtIV complexes is caused by intensive π(O, X) ← eg(Pt) CT transitions.  相似文献   

18.
Structural Investigations on Cs2[B12H12] The crystal structure of Cs2[B12H12] has been determined from X‐ray single‐crystal data collected at room temperature. Dicesium dodecahydro‐closo‐dodecaborate crystallizes as colourless, face‐rich crystals (cubic, Fm 3; a = 1128.12(7) pm; Z = 4). Its synthesis is based on the reaction of Na[BH4] with BF3(O(C2H5)2) via the decomposition of Na[B3H8] in boiling diglyme, followed by subsequent separations, precipitations (with aqueous CsOH solution) and recrystallizations. The crystal structure is best described as anti‐CaF2‐type arrangement with the Cs+ cations in all tetrahedral interstices of the cubic closest‐packed host lattice of the icosahedral [B12H12]2–‐cluster dianions. The intramolecular bond lengths are in the range usually found in closo‐hydroborates: 178 pm for the B–B and 112 pm for the B–H distance. Twelve hydrogen atoms belonging to four [B12H12]2– icosahedra provide an almost perfect cuboctahedral coordination sphere to the Cs+ cations, and their distance of 313 pm (12 ×) attests for the salt‐like character of Cs2[B12H12] according to {(Cs+)2([B12H12]2–)}. The 11B{1H}‐NMR data in aqueous (D2O) solution are δ = –12,70 ppm (1JB–H = 125 Hz), and δ = –15,7 ppm (linewidth: δν1/2 = 295 Hz) for the solid state 11B‐MAS‐NMR.  相似文献   

19.
The protic anions [H(B12X12)] (X: F, Cl, Br, I) are investigated with respect to their ability to protonate neutral molecules in the gas phase by using a combination of tandem mass spectrometry and quantum‐chemical calculations.  相似文献   

20.
Three new series of mixed-ligand clusters of the [(M6X12)X2(RCN)4] (M=Nb, Ta; X=Cl, Br; R=Et, n-Pr, n-Bu) composition have been prepared. It is supposed that four nitrile molecules and two halogen atoms are coordinated to the terminal octahedral coordination sites of the [M6X12]2+ unit.  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