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.
Aluminum-pillared montmorillonite clay, and partially cation-exchanged L type zeolites, efficiently catalyze the selective para-chlorination of phenol using SO2Cl2 in 2,2,4-trimethylpentane at 25 °C. A conversion of ∼96%, a para-selectivity of ∼89%, and a para/ortho ratio of 8.0, were achieved with H+, Al3+, Na+, and K+-L zeolite. This heterogeneous zeolite-catalyzed process is the first example, which shows a pronounced shape-selectivity in the chlorination of phenol.  相似文献   

2.
3.
The crystal structures of the two new synthetic compounds Co2TeO3Cl2 and Co2TeO3Br2 are described together with their magnetic properties. Co2TeO3Cl2 crystallize in the monoclinic space group P21/m with unit cell parameters a=5.0472(6) Å, b=6.6325(9) Å, c=8.3452(10) Å, β=105.43(1)°, Z=2. Co2TeO3Br2 crystallize in the orthorhombic space group Pccn with unit cell parameters a=10.5180(7) Å, b=15.8629(9) Å, c=7.7732(5) Å, Z=8. The crystal structures were solved from single crystal data, R=0.0328 and 0.0412, respectively. Both compounds are layered with only weak interactions in between the layers. The compound Co2TeO3Cl2 has [CoO4Cl2] and [CoO3Cl3] octahedra while Co2TeO3Br2 has [CoO2Br2] tetrahedra and [CoO4Br2] octahedra. The Te(IV) atoms are tetrahedrally [TeO3E] coordinated in both compounds taking the 5s2 lone electron pair E into account. The magnetic properties of the compounds are characterized predominantly by long-range antiferromagnetic ordering below 30 K.  相似文献   

4.
Two new isostructural cobalt selenite halides Co5(SeO3)4Cl2 and Co5(SeO3)4Br2 have been synthesized. They crystallize in the triclinic system space group P−1 with the following lattice parameters for Co5(SeO3)4Cl2: a=6.4935(8) Å, b=7.7288(8) Å, c=7.7443(10) Å, α=66.051(11)°, β=73.610(11)°, γ=81.268(9)°, and Z=1. The crystal structures were solved from single-crystal X-ray data, R1=3.73 and 4.03 for Co5(SeO3)4Cl2 and Co5(SeO3)4Br2, respectively. The new compounds are isostructural to Ni5(SeO3)4Br2.Magnetic susceptibility measurements on oriented single-crystalline samples show anisotropic response in a broad temperature range. The anisotropic susceptibility is quantitatively interpreted within the zero-field splitting schemes for Co2+ and Ni2+ ions. Sharp low-temperature susceptibility features, at TN=18 and 20 K for Co5(SeO3)4Cl2 and Co5(SeO3)4Br2, respectively, are ascribed to antiferromagnetic ordering in a minority magnetic subsystem. In isostructural Ni5(SeO3)4Br2 magnetically ordered subsystem represents a majority fraction (TN=46 K). Nevertheless, anisotropic susceptibility of Ni5(SeO3)4Br2 is dominated at low temperatures by a minority fraction, subject to single-ion anisotropy effects and increasing population of Sz=0 (singlet) ground state of octahedrally coordinated Ni2+.  相似文献   

5.
The crystal structures of 11 synthetic Na-Ca-sulfate-apatites, Na6.45Ca3.55(SO4)6(FxCl1−x)1.55 with x=1 to 0, were refined using X-ray diffraction data yielding residuals between R1=0.0409 and 0.0629 in space group P63/m (Z=1). Lattice constants vary between 9.436(2) and 9.5423(1) Å (for a) and 6.919(2) to 6.8429(1) Å (for c). The sulfate tetrahedra and the two symmetrically independent cation polyhedra about M1 and M2 (occupied by Na and Ca, respectively) are generally very similar to the analogous polyhedra in phosphate apatites. A common structural feature of all members of the solid solution series is a deficiency in the total Cl- and F-content compared with the phosphate apatites. The mean value of (Cl+F) for the solid solution equals 1.55(6) atoms per unit cell compared with the ideal value of 2 atoms per unit cell observed in phosphate apatites. The solid solution series Na6.45Ca3.55(SO4)6Cl1.55-Na6.45 Ca3.55(SO4)6F1.55 shows a gap toward the side of fluoride rich compounds. Under ambient pressure the gap exists between 0<nCl/nCl+nF<0.33, where nCl and nF represent the numbers of Cl- and F-atoms per unit cell.  相似文献   

6.
The mechanism of reaction Cl2+2HBr=2HCl+Br2 has been carefully investigated with density functional theory (DFT) at B3LYP/6-311G** level. A series of three-centred and four-centred transition states have been obtained. The activation energy (138.96 and 147.24 kJ/mol, respectively) of two bimolecular elementary reactions Cl2+HBr→HCl+BrCl and BrCl+HBr→HCl+Br2 is smaller than the dissociation energy of Cl2, HBr and BrCl, indicating that it is favorable for the title reaction occurring in the bimolecular form. The reaction has been applied to the chemical engineering process of recycling Br2 from HBr. Gaseous Cl2 directly reacts with HBr gas, which produces gaseous mixtures containing Br2, and liquid Br2 and HCl are obtained by cooling the mixtures and further separated by absorption with CCl4. The recovery percentage of Br2 is more than 96%, and the Cl2 remaining in liquid Br2 is less than 3.0%. The paper provides a good example of solving the difficult problem in chemical engineering with basic theory.  相似文献   

7.
8.
The complex formed by the reaction of the uranyl ion, UO22+, with bromide ions in the ionic liquids 1-butyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide ([Bmim][Tf2N]) and methyl-tributylammonium bis(trifluoromethylsulfonyl)imide ([MeBu3N][Tf2N]) has been investigated by UV–Vis and U LIII-edge EXAFS spectroscopy and compared to the crystal structure of [Bmim]2[UO2Br4]. The solid state reveals a classical tetragonal bipyramid geometry for [UO2Br4]2− with hydrogen bonds between the Bmim+ and the coordinated bromides. The UV–Vis spectroscopy reveals the quantitative formation of [UO2Br4]2− when a stoichiometric amount of bromide ions is added to UO2(CF3SO3)2 in both Tf2N-based ionic liquids. The absorption spectrum also suggests a D4h symmetry for [UO2Br4]2− in ionic liquids, as previously observed for the [UO2Cl4]2− congener. EXAFS analysis supports this conclusion and demonstrates that the [UO2Br4]2− coordination polyhedron is maintained in the ionic liquids without any coordinating solvent or water molecules. The mean U–O and U–Br distances in the solutions, determined by EXAFS, are, respectively, 1.766(2) and 2.821(2) Å in [Bmim][Tf2N], and, respectively, 1.768(2) and 2.827(2) Å, in [MeBu3N][Tf2N]. Similar results are obtained in both ionic liquids indicating no significant influence of the ionic liquid cation either on the complexation reaction or on the structure of the uranyl species.  相似文献   

9.
A H2O2-HBr system and N-bromosuccinimide in an aqueous medium were used as a ‘green’ approach to electrophilic and radical bromination. Several activated and less activated aromatic molecules, phenylsubstituted ketones and styrene were efficiently brominated ‘on water’ using both systems at ambient temperature and without an added metal or acid catalyst, whereas various non-activated toluenes were functionalized at the benzyl position in the presence of visible light as a radical activator. A comparison of reactivity and selectivity of both brominating systems reveals the H2O2-HBr system to be more reactive than NBS for benzyl bromination and for the bromination of ketones, while for electrophilic aromatic substitution of methoxy-substituted tetralone it was higher for NBS. Also, higher yields of brominated aromatics were observed when using H2O2-HBr ‘on water’. Bromination of styrene reveals that not just the structure of the brominating reagent but the reaction conditions: amount of water, organic solvent, stirring rate and interface structure, play a key role in defining the outcome of bromination (dibromination vs bromohydroxylation). In addition, mild reaction conditions, a straightforward isolation procedure, inexpensive reagents and a lower environment impact make aqueous brominating methods a possible alternative to other reported brominating protocols.  相似文献   

10.
Yttrium- and rare-earth-substituted derivatives of Ca3−vRvCo2O6 (RY, Pr, Nd, Sm, Eu, Gd, Tb, Dy, Ho, Yb, and Lu) have been synthesized and structurally characterized by powder X-ray and neutron diffraction. All phases adopt the K4CdCl6-type structure with space group Rc), in which the trivalent R3+ substituents randomly occupy the Ca2+ site. The homogeneity range of Ca3−vRvCo2O6 extends to v≈0.90 for the substituents concerned. A significant increase in the Co2-O distances within the trigonal-prismatic Co2O6 co-ordination polyhedra upon introduction of R3+ confirms that extra electrons from the R3+-for-Ca2+ substitution exclusively enter the Co2 site of the quasi-one-dimensional Ca3−vRvCo2O6 structure, thereby formally reducing its oxidation state. This is furthermore supported by magnetic susceptibility and low-temperature neutron diffraction measurements. The long-range ferrimagnetic ordering temperature is reduced upon R substitution and appears to vanish for v>∼0.30.  相似文献   

11.
The luminescence properties of Cs3Bi2Cl9, α-Cs3Sb2Cl9, and β-Cs3Sb2Cl9 are reported and compared with those of Cs3Bi2Br9. The first two compounds have comparable luminescence properties which can be described in terms of a band model. Deep center emission is observed for both compounds, whereas edge emission is observed only for Cs3Bi2Cl9. The optical transitions of β-Cs3Sb2Cl9 are localized on the Sb3+ ion. The orientation of the lone-pair orbitals of the ns2 ions seems to play an important role in the formation of the cationic valence band. The α-β transformation must therefore have a considerable influence on the spectral properties of Cs3Sb2Cl9.  相似文献   

12.
A re-interpretation and re-evaluation of single-crystal X-ray diffraction data of a previously reported ‘(NH4)2(NH3)[Ni(NH3)2Cl4]’ (J. Solid State Chem. 162 (2001) 254) give a new formula (NH4)2−2z[Ni(NH3)2]z[Ni(NH3)2Cl4] with z=0.152. This new formula results from defects in an idealized ‘(NH4)2[Ni(NH3)2Cl4]’ basic structure, where two adjacent NH4+ cations are replaced by one Ni(NH3)22+ unit. Cl anions from the basic structure complete the coordination sphere of the new Ni2+ to [Ni(NH3)2Cl4]2−.  相似文献   

13.
Microcrystalline samples of Zn(NH3)2Br2 and Ni(NH3)2X2 (X is Cl and Br) have been investigated from 100 to 293 K using X-ray diffraction and IR spectroscopy measurements (range 400–4000 cm) performed with isotopically dilute (5% deuterated) samples. Values of Δν(ND)/ΔT for all compounds hint at the existence of hydrogen bonds. Zn(NH3)2Br2 shows The dynamics of ammonia molecules even at 100 K, and no indications are apparent that dynamic disorder of ammonia molecules takes place in Ni(NH3)2X2 (X is Cl and Br). A comparison between octahedrally coordinated ammoniates [Ni(NH3)6]Br2, Ni(NH3)2Br2 and [Zn(NH3)6]Br2 with tetrahedrally coordinated ones [Zn(NH3)2Br2] leads to the conclusion that the lower coordination number increases the strength of the hydrogen bonds. Because this effect is small, it is not possible to separate the influence of the type of coordinating ions for one coordination number from the influence of the coordination number itself.  相似文献   

14.
Arrested relaxation infrared chemiluminescence studies of the H + Cl2, SCl2, S2Cl2, SOCl2 and SO2-Cl2 reactions have been made. The mean fraction of vibrational (stational) energy released to HCl is 0.40 (0.10); 0.40 (0.13); 0.38 (? 0.02); 0.33 (?0.02) and 0.36 (?0.02) for the series. Only the H + SCl2 reaction shows a two component initial rotational distribution. The larger (fV) and (fR) from H + SCl2, relative to the other polyatomic reagents, is consistent with the observation that this is the only reaction that shows forward scattering. The room temperature rate constants also were measured, relative to H + Cl2, and were found to decline in the series from 0.68 for SCl2 to 0.02 for SO2Cl2. All of these data support the suggestion (first made by Heydtmann and Polanyi) that the unusual rotational energy disposal pattern from H + SCl2 is a consequence of migration of H from the initially encountered C1 to the second C1, which then forms the HCI product; this pathway augments the direct reactive pathway, which gives HCI in lower J states.  相似文献   

15.
The crystal structure of Ca12Al14O32Cl2 was determined from laboratory X-ray powder diffraction data (CuKα1) using the Rietveld method, with the anisotropic displacement parameters being assigned for all atoms. The crystal structure is cubic (space group , Z=2) with lattice dimensions a=1.200950(5) nm and V=1.73211(1) nm3. The reliability indices calculated from the Rietveld method were Rwp=8.48% (S=1.21), Rp=6.05%, RB=1.27% and RF=1.01%. The validity of the structural model was verified by the three-dimensional electron density distribution, the structural bias of which was reduced as much as possible using the maximum-entropy methods-based pattern fitting (MPF). The reliability indices calculated from the MPF were RB=0.75% and RF=0.56%. In the structural model there are one Ca site, two Al sites, two O sites and one Cl site. This compound is isomorphous with Ca12Al10.6Si3.4O32Cl5.4. Europium-doped sample Ca12Al14O32Cl2:Eu2+ was prepared and the photoluminescence properties were presented. The excitation spectrum consisted of two wide bands, which were located at about 268 and 324 nm. The emission spectrum, when excited at 324 nm, resulted in indigo light with a peak at about 442 nm.  相似文献   

16.
Adducts of cucurbit[6]uril with Ca2+ and trinuclear cluster chloroaquacomplexes (H9O4)2(H7O3)2[(Ca(H2O)5)2(C36H36N24O12)]Cl8·0.67H2O (1) and [(Ca(H2O)5)2(C36H36N24O12)]× [Mo3O2S2Cl6(H2O)3]2·13H2O (2) are obtained and structurally characterized. The structures of both compounds contain polymeric [Ca(H2O) n ]22 CB[6]∞ cations that form infinite columns; the space between them is filled with Cls- (1) and [Mo3O2S2Cl6(H2O)3]2s- (2). A new (H7O3)2(H5O2)× [Mo3S4Cl6.25Br0.25(H2O)2](C36H36N24O12)·CH2Cl2·6H2O complex (3) is also obtained and structurally characterized.  相似文献   

17.
The new compound Nb3Se5Cl7 was prepared by heating 2NbSe2Cl2 + 1NbCl4 at 530°C for 2–3 weeks. The compound is monoclinic with a = 7.599, b = 12.675, c = 8.051Å; β = 106.27°; space group P21m. The corresponding bromide, Nb3Se5Br7 (obtained by decomposition of NbSe2Br2 under NbSeBr3), is isotypic with a = 7.621, b = 12.833, c = 8.069Å; β = 106.21°. from the crystal structure and XPS spectra it follows that Nb3Se5Cl7 can be formulated as: [Nb4+2Nb5+1(Se2)2?2Se2?1Cl?7]. The structure consists of chains of composition [Nb4+2(Se2)2?2Cl?5], to which side chains [Nb5+Se2?Cl?2] are attached. The Nb4+ atoms form pairs (NbNb = 2.94 Å) which explains that Nb3Se5Cl7 is a diamagnetic semiconductor with a band gap (1.59 eV at 5°K, 1.49 eV at 300°K) very similar to that of NbSe2Cl2.  相似文献   

18.
The present work is dedicated to the XRD, ED and HREM characterization of a new bismuth copper oxyphosphate Bi∼6.2Cu∼6.2O8(PO4)5 (a=11.599(2)Å, , c=37.541(5)Å, R1=0.0755, Rw2=0.174, G.S Pn21a). The relatively long size of its c parameter is due to the arrangement along this direction of two kinds of ribbon-like polycations formed by edge sharing O(Bi, Cu)4 tetrahedra. The existence of such cations is characterized by the b∼5.2 Å value intrinsic to the ribbons structure and commonly found in bismuth oxyphosphate materials. In the title compound, 2-tetrahedra wide [Bi∼2.4Cu∼3.6O4]6.4+ and 3-tetrahedra wide [Bi∼5Cu∼3O6]9+ ribbons are isolated by phosphate groups and alternate along c. The interstitial site created between two different sizes ribbons is occupied by Cu2+ cations disordered over several close crystallographic sites. The mixed Bi3+/Cu2+ nature of certain edge-of-ribbons positions induces a disorder over several configurations of the phosphate groups. The concerned oxygen atoms form the environment of the disordered interstitial Cu2+ cations which occupy tunnels formed by the phosphate anions. The high-resolution electron microscope study enables a precise correlation between the observed images and the refined crystal structure, evidencing the polycations visualization. Furthermore, this material being the second example of partially disordered compound similar chemical system, some topological rules can be deduced. The b-axis doubling was observed by ED and HREM and is assigned to the ordering of interstitial Cu2+ within tunnels cations. A partial intra-tunnel ordering was also observed.  相似文献   

19.
The sections Li2MCl4?4xBr4x of the quaternary systems LiCl-LiBr-MCl2-MBr2 with M = Mn, Cd, and Fe were studied by high-temperature X-ray diffraction patterns and DTA and DSC measurements. In the quasibinary lithium manganese halide system complete series of solid solutions exist between the inverse spinels Li2MnCl4 and Li2MnBr4. Li2MnBr4 and solid solutions with x > 0.54 undergo phase transitions to tetragonal spinels at lower temperatures. In the nonquasibinary system with M = Cd, only at temperatures near 400°C a complete series of mixed crystals is formed. At lower temperatures the system is mainly two-phase with rock salt-type Li1?yCd0.5yCl1?xBrx and cadmium chloride-type Cd1?yLi2yCl2?2xBr2x solid solutions in equilibrium. The lithium iron halide system is similar to that of cadmium, but spinel-type Li2FeBr4 does not exist at any temperature. The manganese and cadmium halide spinels and spinel solid solutions undergo phase transitions to NaCl defect structures at elevated temperatures.  相似文献   

20.
The crystal structure of the new Bi∼3Cd∼3.72Co∼1.28O5(PO4)3 has been refined from single crystal XRD data, R1=5.37%, space group Abmm, a=11.5322(28) Å, b=5.4760(13) Å, c=23.2446(56) Å, Z=4. Compared to Bi∼1.2M∼1.2O1.5(PO4) and Bi∼6.2Cu∼6.2O8(PO4)5, this compound is an additional example of disordered Bi3+/M2+ oxyphosphate and is well described from the arrangement of double [Bi4Cd4O6]8+ (=D) and triple [Bi2Cd3.44Co0.56O4]6+ (=T) polycationic ribbons formed of edge-sharing O(Bi,M)4 tetrahedra surrounded by PO4 groups. According to the nomenclature defined in this work, the sequence is TT/DtDt, where t stands for the tunnels created by PO4 between two subsequent double ribbons and occupied by Co2+. The HREM study allows a clear visualization of the announced sequence by comparison with the refined crystal structure. The Bi3+/M2+ statistic disorder at the edges of T and D entities is responsible for the PO4 multi-configuration disorder around a central P atom. Infrared spectroscopy and neutron diffraction of similar compounds (without the highly absorbing Cadmium) even suggests the long range ordering loss for phosphates. Therefore, electron diffraction shows the existence of a modulation vector q*=1/2a*+(1/3+ε)b* which pictures cationic ordering in the (001) plane, at the crystallite scale. This ordering is largely lost at the single crystal scale. The existence of mixed Bi3+/M2+ positions also enables a partial filling of the tunnels by Co2+ and yields a composition range checked by solid state reaction. The title compound can be prepared as a single phase and also the M=Zn2+ term can be obtained in a biphasic mixture. For M=Cu2+, a monoclinic distortion has been evidenced from XRD and HREM patterns but surprisingly, the orthorhombic ideal form can also be obtained in similar conditions.  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