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 15 毫秒
1.
Preparation and Spectroscopic Characterization of Nonahalogenodiiridates(III), [Ir2X9]3?, X = Cl, Br The pure nonahalogenodiiridates(III), A3[Ir2X9] (A = K, Cs, tetraalkylammonium; X = Cl, Br) have been prepared. They are formed from the monomer hexahalogenoiridates(III) which are bridged to confacial bioctahedral complexes by ligand abstraction in less polar organic solvents. The IR and Raman spectra exhibit bands in three characteristic regions; at high wavenumbers stretching vibrations with terminal ligands ν(Ir?Clt): 360?300, ν(Ir?Brt): 250?220; in a middle region with bridging ligands ν(Ir?Clb): 290?235, ν(Ir?Brb): 205?190 cm?1; the deformation bands are observed at distinct lower frequencies. The distance between ν(Ir?Xt) and ν(Ir?Xb) increases with decreasing size of the cations. The electronic spectra measured at thin films of the pure complex salts at 10 K show some intensive charge transfer transitions in the UV and one or two weak d? d bands in the visible region.  相似文献   

2.
3.
Thermodynamics and Kinetics of the Xa-Substitution of [W6Cl8]X6a(2?) and [Mo6Cl8]X6a(2?) Complexes; (X = Cl, Br, I) The title subject has been investigated in different solvent mixtures (see “Inhaltsübersicht”). In some cases the progress of the reaction has been followed by an emf method; in most cases the reaction was stopped after definite times by precipitation of the oxiniumsalts. Thermodynamics. For equilibria of the type (a) one finds the constant \documentclass{article}\pagestyle{empty}\begin{document}$ {\rm C} = \frac{{[{\rm Br}^{\rm a} ][{\rm Cl}^ - ]}}{{[{\rm Cl}^{\rm a} ][{\rm Br}^ - ]}} $\end{document}, where [Bra] and [Cla] mark the total number of Br or Cl occupying Xa-positions of the complex. The Xa-positions are thermodynamically equivalent, the substitution proceeds statistically, so that the steps of reaction (a) with the equilibrium constants K1 to K6 are given by \documentclass{article}\pagestyle{empty}\begin{document}$ {\rm K} = \frac{{{\rm W}({\rm hin})}}{{{\rm W}({\rm r\ddot uck})}} \cdot {\rm C} $\end{document} if W(hin) and W(rück) describe the probability of the forward and the back reaction. Similarly in some simple complexes (e. g. Irx62?);PdX42? the statistical effect plays a dominating role. The kinetics may be described as (b) The aquotation step is rate determining. Consequently the reaction of the first order. Rate constants for the forward and the reversed reaction between 0 and 25°C have been measured. The activation energy is ≈ 18 kcal. With the molybdenum complexes the Xa-substitutions is about 10 times faster as with the tungsten complexes.  相似文献   

4.
Ligand-Exchange Reactions on TcNX4? Complexes (X = Cl, Br) Ligand exchange reactions on the nitridotechnetium(VI) compounds (Bu4N)TcNCl4 and (Bu4N)TcNBr4 are reported. The use of various organic ligands having different donor atom sets produces TcV nitrido complexes. The reaction of (Bu4N)TcNCl4 with (Bu4N)TcNBr4 is characterized by the formation of TcVI complex species with mixed Cl/Br coordination spheres. EPR detection of the mixed-ligand complexes results in a well-defined dependence of the EPR parameters on the composition of the first coordination sphere of the complexes.  相似文献   

5.
Preparation and Characterization of [Pt(mal)2]2? and trans-[Pt(mal)2X2]2? (X = Cl, Br, I, SCN) By twofold treatment of K2[PtCl4] with potassium hydrogen malonate in a queous solution the yellow K2[Pt(mal)2] · H2O is obtained. After extraction with tetrabutylammonium ions into dichloromethane by oxidative addition at ?90°C the PtIV complexes [Pt(mal)2X2]2?, X = Cl, Br, I, SCN, are formed. The SCN ligands are coordinated to Pt via S. The IR and Raman spectra are discussed and assigned.  相似文献   

6.
The perhalogenated closo‐dodecaborate dianions [B12X12]2? (X=H, F, Cl, Br, I) are three‐dimensional counterparts to the two‐dimensional aromatics C6X6 (X=H, F, Cl, Br, I). Whereas oxidation of the parent compounds [B12H12]2? and benzene does not lead to isolable radicals, the perhalogenated analogues can be oxidized by chemical or electrochemical methods to give stable radicals. The chemical oxidation of the closo‐dodecaborate dianions [B12X12]2? with the strong oxidizer AsF5 in liquid sulfur dioxide (lSO2) yielded the corresponding radical anions [B12X12] ? ? (X=F, Cl, Br). The presence of radical ions was proven by EPR and UV/Vis spectroscopy and supported by quantum chemical calculations. Use of an excess amount of the oxidizing agent allowed the synthesis of the neutral perhalogenated hypercloso‐boranes B12X12 (X=Cl, Br). These compounds were characterized by single‐crystal X‐ray diffraction of dark blue B12Cl12 and [Na(SO2)6][B12Br12] ? B12Br12. Sublimation of the crude reaction products that contained B12X12 (X=Cl, Br) resulted in pure dark blue B12Cl12 or decomposition to red B9Br9, respectively. The energetics of the oxidation processes in the gas phase were calculated by DFT methods at the PBE0/def2‐TZVPP level of theory. They revealed the trend of increasing ionization potentials of the [B12X12]2? dianions by going from fluorine to bromine as halogen substituent. The oxidation of all [B12X12]2? dianions was also studied in the gas phase by mass spectrometry in an ion trap. The electrochemical oxidation of the closo‐dodecaborate dianions [B12X12]2? (X=F, Cl, Br, I) by cyclic and Osteryoung square‐wave voltammetry in liquid sulfur dioxide or acetonitrile showed very good agreement with quantum chemical calculations in the gas phase. For [B12X12]2? (X=F, Cl, Br) the first and second oxidation processes are detected. Whereas the first process is quasi‐reversible (with oxidation potentials in the range between +1.68 and +2.29 V (lSO2, versus ferrocene/ferrocenium (Fc0/+))), the second process is irreversible (with oxidation potentials ranging from +2.63 to +2.71 V (lSO2, versus Fc0/+)). [B12I12]2? showed a complex oxidation behavior in cyclic voltammetry experiments, presumably owing to decomposition of the cluster anion under release of iodide, which also explains the failure to isolate the respective radical by chemical oxidation.  相似文献   

7.
Preparation and spectroscopic characterization of the decahalogenodirhenates(IV), [Re2X10]2?, X = Cl, Br On heating of [ReX6]2? with trifluoroacetic acid/trifluoroacetic anhydride (1 : 1), the edge-sharing bioctahedral anions [Re2X10]2?, X = Cl, Br are formed, which IR and Raman spectra are assigned according to point group D2h. The bands are found in three characteristic regions; at high wavenumbers stretching vibrations with terminal ligands v(ReClt): 367–321, v(ReBrt): 242–195; in an intermediate region with bridging ligands v(ReClb): 278–250, v(ReBrb): 201–167 cm?1, and at distinct lower frequencies the deformation modes. The absorption spectra of the dirhenates are distinguished in the region 600–1400 nm by eight intraconfigurational transitions with a slight bathochromic shift and higher intensities in comparison to the monomeric complexes. Due to a stronger bonding of the terminal ligands the energy of the charge transfer bands is lowered by about 4 000 cm?1, too. The magnetic moments are 3.32 and 3.81 B.M./ReIV for [Re2Cl10]2? and [Re2Br10]2?, respectively.  相似文献   

8.
The title compound, 3,3′‐(4‐pyridyl­imino)­di­propane­nitrile, C11H12N4, has a twofold axis and consists of a pyridine ring head and two cyano­ethyl tails, the three groups being linked by an N atom. The planar geometry around the amino N atom suggests conjugation with the π‐system of the pyridine ring. The mol­ecules are stacked in a layer structure via relatively weak to very weak intermolecular C—H⃛π and C—H⃛N hydrogen‐bond interactions.  相似文献   

9.
Polynuclear Complexes with Fe? As, Fe? Sb, and Fe? Bi Frameworks The anionic iron clusters Fe3(CO)112? and Fe4(CO)132? were reacted with compounds EX3 and with organic derivatives REX2 and R2EX of the elements arsenic, antimony, and bismuth. Commonly redox and cluster degradation reactions were observed. The new complexes [(CO)4Fe? AsMe2? Fe(CO)4]?, [HFe3(CO)9(mu;3-SbBut)]?, [Fe3(CO)10 (mu;3-Sb)]?, and [Fe3(CO)10(mu;3-Bi)]? were formed and isolated as their PPN salts. The Fe? As? Fe complex was identified by a structure determination, the other complexes were identified by their spectra.  相似文献   

10.
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   

11.
Ruthenium(III) Phthalocyanines: Synthesis and Properties of Di(halo)phthalocyaninato(1?)ruthenium(III) Di(halo)phthalocyaninato(1?)ruthenium(III), [Ru(X)2Pc?] (X = Cl, Br, I) is prepared by oxidation of [Ru(X)2Pc2?]? (Cl, Br, OH) with halogene in dichloromethane. The magnetic moment of [Ru(X)2Pc?] is 2,48 μB (X = Cl) resp. 2,56 μB (X = Br) in accordance with a systeme of two independent spins (low spin RuIII and Pc?: S = 1/2). The optical spectra of the red violet solution of [Ru(X)2Pc?] (Cl, Br) are typical for the Pc? ligand with the “B” at 13.5 kK, “Q1” at 19.3 kK and “Q2 region” at 31.9 kK. Sytematic spectral changes within the iron group are discussed. The presence of the Pc? ligand is confirmed by the vibrational spectra, too. Characteristic are the metal dependent bands in the m.i.r. spectra at 1 352 and 1 458 cm?1 and the strong Raman line at 1 600 cm?1. The antisymmetric Ru? X stretch (vas(Ru? X)) is observed at 189 cm?1 (X = I) resp. 234 cm?1 (X = Br). There are two interdependent bands at 295 and 327 cm?1 in the region expected for vas(Ru? Cl) attributed to strong interaction of vas(Ru? Cl) with an out-of-plane Pc? tilting mode of the same irreducible representation. Only the symmetric Ru? Br stretch at 183 cm?1 is selectively enhanced in the resonance-Raman(RR) spectra. The Raman line at 168 cm?1 of the diiodo complex is assigned to loosely bound iodine. The broad band at 978 cm?1 in the RR spectra of the dichloro complex is due to an intraconfigurational transition within the electronic ground state of low spin RuIII split by spin orbit coupling.  相似文献   

12.
Preparation and Spectroscopic Characterization of the Fluorophosphonium Salts X2FPSCH3+MF6? (X = Br, Cl; M = As, Sb) and XF2PSCH3+SbF6? (X = Br, Cl, F) The preparation of the fluorophosphonium salts X2FPSCH3+MF6? (X = Br, Cl; M = As, Sb) and XF2PSCH3+SbF6? (X = Br, Cl, F) by methylation of the corresponding thiophosphorylhalides in the system CH3F/SO2/MF5 (M = As, Sb) is reported. The new salts are characterized by their vibrational and NMR spectra.  相似文献   

13.
The structures of seven A2Cu4X10 compounds containing quasi‐planar oligomers are reported: bis(1,2,4‐trimethylpyridinium) hexa‐μ‐chlorido‐tetrachloridotetracuprate(II), (C8H12N)2[Cu4Cl10], (I), and the hexa‐μ‐bromido‐tetrabromidotetracuprate(II) salts of 1,2,4‐trimethylpyridinium, (C8H12N)2[Cu4Br10], (II), 3,4‐dimethylpyridinium, (C7H10N)2[Cu4Br10], (III), 2,3‐dimethylpyridinium, (C7H10N)2[Cu4Br10], (IV), 1‐methylpyridinium, (C6H8N)2[Cu4Br10], (V), trimethylphenylammonium, (C9H14N)2[Cu4Br10], (VI), and 2,4‐dimethylpyridinium, (C7H10N)2[Cu4Br10], (VII). The first four are isomorphous and contain stacks of tetracopper oligomers aggregated through semicoordinate Cu...X bond formation in a 4(,) stacking pattern. The 1‐methylpyridinium salt also contains oligomers stacked in a 4(,) pattern, but is isomorphous with the known chloride analog instead. The trimethylphenylammonium salt contains stacks of oligomers arranged in a 4(,) stacking pattern similar to the tetramethylphosphonium analog. These six structures feature inversion‐related organic cation pairs and hybrid oligomer/organic cation layers derived from the parent CuX2 structure. The 2,4‐dimethylpyridinium salt is isomorphous with the known (2‐amino‐4‐methylpyridinium)2Cu4Cl10 structure, in which isolated stacks of organic cations and of oligomers in a 4(,) pattern are found. In bis(3‐chloro‐1‐methylpyridinium) octa‐μ‐bromido‐tetrabromidopentacuprate(II), (C6H7ClN)[Cu5Br12], (VIII), containing the first reported fully halogenated quasi‐planar pentacopper oligomer, the oligomers stack in a 5(,) stacking pattern as the highest nuclearity [CunX2n+2]2− oligomer compound known with isolated stacking. Bis(2‐chloro‐1‐methylpyridinium) dodeca‐μ‐bromido‐tetrabromidoheptacuprate(II), (C6H7ClN)2[Cu7Br16], (IX), contains the second heptacopper oligomer reported and consists of layers of interleaved oligomer stacks with a 7[(,)][(−,−)] pattern isomorphous with that of the known 1,2‐dimethylpyridinium analog. All the oligomers reported here are inversion symmetric.  相似文献   

14.
Synthesis and Spectroscopical Characterization of Di(halo)phthalocyaninato(1–)rhodium(III), [RhX2Pc1?] (X = Cl, Br, I) Bronze-coloured di(halo)phthalocyaninato(1–)-rhodium(III), [RhX2Pc1?] (X = Cl, Br) and [RhI2Pc1?] · I2 is prepared by oxidation of (nBu4N)[RhX2Pc2?] with the corresponding halogene. Irrespective of the halo ligands, two irreversible electrode reactions due to the first ringreduction (ER = ?0,90 V) and ringoxidation (EO = 0,82 V) are present in the cyclovoltammogram of (nBu4N)[RhX2Pc2?]. The optical spectra show typical absorptions of the Pc1?-ligand at 14.0 kK and 19.1 kK. Characteristic vibrational bands are at 1 366/1 449 cm?1 (i. r.) and 569/1 132/1 180/1 600 cm?1 (resonance Raman (r. r.)). The antisym. (Rh? X)-stretching vibration is observed at 294 cm?1 (X = Cl), 240 cm?4 (Br) and 200 cm?1 (I). Only the sym. (Rh? I)-stretching vibration at 133 cm?1 is r. r. enhanced together with a strong line at 170 cm?1, which is assigned to the (I? I)-stretching vibration of the incorporated iodine molecule. Both modes show overtones and combinationbands.  相似文献   

15.
The aim of this study is the determination of the g tensor of the tetrathiafulvalene (TTF) molecule involved in chlorine and bromine radical–ion salts. This work is based on ab initio calculations using several basis sets which enabled us to compare theoretical and experimental measurement data. The results show clearly the impact of the structural distortions on the g gyroscopic matrix elements and proves the important fact that even a small variation of the crystallographic parameters has major consequences on the physical–chemical properties. © 2002 John Wiley & Sons, Inc. Int J Quantum Chem, 2002  相似文献   

16.
Specific short contacts are important in crystal engineering. Hydrogen bonds have been particularly successful and together with halogen bonds can be useful for assembling small molecules or ions into crystals. The ionic constituents in the isomorphous 3,5‐dichloropyridinium (3,5‐diClPy) tetrahalometallates 3,5‐dichloropyridinium tetrachloridozincate(II), (C5H4Cl2N)2[ZnCl4] or (3,5‐diClPy)2ZnCl4, 3,5‐dichloropyridinium tetrabromidozincate(II), (C5H4Cl2N)2[ZnBr4] or (3,5‐diClPy)2ZnBr4, and 3,5‐dichloropyridinium tetrabromidocobaltate(II), (C5H4Cl2N)2[CoBr4] or (3,5‐diClPy)2CoBr4, arrange according to favourable electrostatic interactions. Cations are preferably surrounded by anions and vice versa ; rare cation–cation contacts are associated with an antiparallel dipole orientation. N—H…X (X = Cl and Br) hydrogen bonds and X X halogen bonds compete as closest contacts between neighbouring residues. The former dominate in the title compounds; the four symmetrically independent pyridinium N—H groups in each compound act as donors in charge‐assisted hydrogen bonds, with halogen ligands and the tetrahedral metallate anions as acceptors. The M X coordinative bonds in the latter are significantly longer if the halide ligand is engaged in a classical X …H—N hydrogen bond. In all three solids, triangular halogen‐bond interactions are observed. They might contribute to the stabilization of the structures, but even the shortest interhalogen contacts are only slightly shorter than the sum of the van der Waals radii.  相似文献   

17.
This paper gives an account on hypervalent fluoro‐ and chloro(pentafluoroethyl)germanium compounds. The selective synthesis of the tris(pentafluoroethyl)dichlorogermanate salt [PNP][(C2F5)3GeCl2] as well as its X‐ray structural analysis is described. As a representative example for pentafluoroethylfluorogermanates, the synthesis and structure of 2,4,6‐triphenylpyryliumtris(pentafluoroethyl)difluorogermanate [C23H17O][(C2F5)3GeF2] is reported. Fluoride‐ion affinities for pentafluoroethylgermanes were calculated using quantum chemical methods, disclosing (C2F5)3GeF as a weaker Lewis acid than (C2F5)3SiF or (C2F5)3PF2. The theoretical results were confirmed by experiments and give the basis of a synthetic protocol for (C2F5)3GeF. Pentakis(pentafluoroethyl)germanate [PPh4][Ge(C2F5)5] was detected as an intermediate during the synthesis of [PPh4][(C2F5)4GeF] starting from tris(pentafluoroethyl)difluorogermanate and LiC2F5.  相似文献   

18.
The title compounds, {4,4′‐di­bromo‐2,2′‐[1,3‐propane­diyl­bis(nitrilo­methyl­idyne‐N)]­diphenolato‐O,O′}nickel(II), [Ni(C17­H14­Br2­N2O2)], and {4,4′‐di­chloro‐2,2′‐[1,3‐pro­pane­diyl­bis­(ni­trilo­methyl­idyne‐N)]­di­phen­ol­ato‐O,O′}­copper(II), [Cu­(C17­H14­Cl2­N2O2)], lie on crystallographic twofold axes. In both structures, the metal coordination sphere is a tetrahedrally distorted square plane formed by the four‐coordinate N2O2 donor set of the Schiff base imine–phenol ligands. In the Ni compound, the Ni—O and Ni—N distances are 1.908 (3) and 1.959 (4) Å, respectively, while in the Cu compound, the Cu—O and Cu—N distances are 1.907 (2) and 1.960 (2) Å, respectively. The two Schiff base moieties, which themselves are nearly planar, are inclined at an angle of 29.26 (7)° for the Ni compound and 29.26 (5)° for the Cu compound.  相似文献   

19.
Preparation, Vibrational Spectra and Normal Coordinate Analysis of Decahalogenoditechnetates(IV), [Tc2X10]2?, X = Cl, Br The reaction of [TcX6]2?, X = Cl, Br, with trifluoroacetic acid yield at room temperature the edge-sharing bioctahedral anions [Tc2X10]2?, which IR and Raman spectra are assigned according to point group D2h. Using the crystal data of isostructural osmium complexes a normal coordinate analysis based on a general valence force field has been performed for [Tc2X10]2?, revealing a good agreement of the calculated frequencies with the bands observed in the IR and Raman spectra. The stronger bonding of the terminal as compared to the bridging ligands is shown by the valence force constants, fd(TcXt) > Fd(TcXb).  相似文献   

20.
The sterically encumbered ter­phenyl halides 2′‐chloro‐2,2′′,4,4′′,6,6′′‐hexaisopropyl‐1,1′:3′,1′′‐terphenyl, C36H49Cl, (I), 2′‐bromo‐2,2′′,4,4′′,6,6′′‐hexaisopropyl‐1,1′:3′,1′′‐terphenyl, C36H49Br, (II), and 2′‐iodo‐2,2′′,4,4′′,6,6′′‐hexaisopropyl‐1,1′:3′,1′′‐terphenyl, C36H49I, (III), crystallize in space group Pnma. They are isomorphous and isostructural with a plane of symmetry through the centre of the mol­ecule. The C–halide bond distances are 1.745 (3), 1.910 (4) and 2.102 (6) Å for (I)–(III), respectively.  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