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.
Hierarchical hybridized nanocomposites with rationally constructed compositions and structures have been considered key for achieving superior Li‐ion battery performance owing to their enhanced properties, such as fast lithium ion diffusion, good collection and transport of electrons, and a buffer zone for relieving the large volume variations during cycling processes. Hierarchical MoS2@carbon microspheres (HMCM) have been synthesized in a facile hydrothermal treatment. The structure analyses reveal that ultrathin MoS2 nanoflakes (ca. 2–5 nm) are vertically supported on the surface of carbon nanospheres. The reversible capacity of the HMCM nanocomposite is maintained at 650 mA h g?1 after 300 cycles at 1 A g?1. Furthermore, the capacity can reach 477 mA h g?1 even at a high current density of 4 A g?1. The outstanding electrochemical performance of HMCM is attributed to the synergetic effect between the carbon spheres and the ultrathin MoS2 nanoflakes. Additionally, the carbon matrix can supply conductive networks and prevent the aggregation of layered MoS2 during the charge/discharge process; and ultrathin MoS2 nanoflakes with enlarged surface areas, which can guarantee the flow of the electrolyte, provide more active sites and reduce the diffusion energy barrier of Li+ ions.  相似文献   

2.
Metal Complexes of Functionalized Sulfur‐containing Ligands. XVII Synthesis of S ‐Oxides of 1,2,4‐Trithiolane, 1,2,4,5‐Tetrathiane as well as 1,2,3,5,6‐Pentathiepane, and their Reactions with (Ph3P)2Pt(η2‐C2H4). X‐Ray Structure Analysis of 3,3,5,5‐Tetraphenyl‐1,2,4‐trithiolane 1‐oxide 3,3,5,5‐Tetraphenyl‐1,2,4‐trithiolan ( 1 ) was oxidized using m‐chloroperbenzoic acid to give, selectively, the 3,3,5,5‐tetraphenyl‐1,2,4‐trithiolane 1‐oxide ( 2 ). 2 was characterized structurally. The reaction of octamethyl tetrathiadispiro[3.2.3.2]dodecane‐2,9‐dione ( 3 ) with trifluoroperacetic acid at –50 °C yielded the corresponding 5‐oxide 4 . Oxidation of octamethyl pentathiadispiro[3.3.3.2]tridecane‐2,9‐dione ( 5 ) with m‐chloroperbenzoic acid at 0 °C gave the 12‐oxide 6 . Treatment of 2 with two equivalents of (Ph3P)2Pt(η2‐C2H4) ( 7 ) afforded a mixture (1 : 1) of the complexes (Ph3P)2PtSCPh2S ( 8 ) and (Ph3P)2Pt(η2‐Ph2C=S=O) ( 9 ), respectively.  相似文献   

3.
In the course of our investigations on polymetallic complexes derived from 1,3‐bis(thiophosphinoyl)indene (Ind(Ph2P?S)2), we observed original fluxional behavior and report herein a joint experimental/computational study of this dynamic process. Starting from the indenylidene chloropalladate species [Pd{Ind(Ph2P?S)2}Cl]? ( 1 ), the new PdII???RhI hetero‐bimetallic pincer complex [PdCl{Ind(Ph2P?S)2}Rh(nbd)] ( 2 ; nbd=2,5‐norbornadiene) was prepared. X‐ray crystallography and DFT calculations substantiate the presence of a d8???d8 interaction. According to multinuclear variable‐temperature NMR spectroscopic experiments, the pendant {Rh(nbd)} fragment of 2 readily shifts in solution at room temperature between the two edges of the SCS tridentate ligand. To assess the role of the pincer‐based polymetallic structure on this fluxional behavior, the related monometallic Rh complex [Rh{IndH(Ph2P?S)2}(nbd)] ( 3 ) was prepared. No evidence for a metal shift was observed in that case, even at high temperature, thus indicating that inplane pincer coordination to the Pd center plays a crucial role. The previously described PdII???IrI bimetallic complex 4 exhibited fluxional behavior in solution, but with a significantly higher activation barrier than 2 . This finding demonstrates the generality of this metal‐shift process and the strong influence of the involved metal centers on the associated activation barrier. DFT calculations were performed to shed light onto the mechanism of such metal‐shift processes and to identify the factors that influence the associated activation barriers. Significantly different pathways were found for bimetallic complexes 2 and 4 on one hand and the monometallic complex 3 on the other hand. The corresponding activation barriers predicted computationally are in very good agreement with the experimental observations.  相似文献   

4.
The reactions of thiophene‐2‐(N‐diphenylphosphino)methylamine, Ph2PNHCH2‐C4H3S, 1 and thiophene‐2‐[N,N‐bis(diphenylphosphino)methylamine], (Ph2P)2NCH2‐C4H3S, 2, with MCl2(cod) (M = Pd, Pt; cod = 1,5‐cyclooctadiene) or [Cu(CH3CN)4]PF6 yields the new complexes [M(Ph2PNHCH2‐C4H3S)2Cl2], M = Pd 1a, Pt 1b, [Cu(Ph2PNHCH2‐C4H3S)4]PF6, 1c, and [M(Ph2P)2NCH2‐C4H3S)Cl2], M = Pd 2a, Pt 2b, {Cu[(Ph2P)2NCH2‐C4H3S]2}PF6, 2c, respectively. The new compounds were isolated as analytically pure crystalline solids and characterized by 31P‐, 13C‐, 1H‐NMR and IR spectroscopy and elemental analysis. Furthermore, the solid‐state molecular structures of representative palladium and platinum complexes of bis(phosphine)amine, 2a and 2b, respectively, were determined using single crystal X‐ray diffraction analysis. The palladium complexes were tested as potential catalysts in the Heck and Suzuki cross‐coupling reactions. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

5.
The in situ synthesis of ethylene‐co‐norbornene copolymers/multi‐walled carbon nanotubes (MWNTs) nanocomposites was achieved by rare‐earth half‐sandwich scandium precursor [Sc(η5‐C5Me4SiMe3)(η1‐CH2SiMe3)2(THF)] (1) activated by [Ph3C][B(C6F5)4], through a non‐PFT (Polymerization Filling Technique) approach. MWNTs nanocomposites with low aluminum residue were obtained with excellent yields even though small amounts of triisobutylaluminium were needed as scavenger to prevent catalyst poisoning by MWNT impurities. MWNT bundles were disaggregated and highly coated with Poly(ethylene‐co‐norbornene) [P(E‐co‐N)] as revealed by transmission electron microscopy. Interestingly, P(E‐co‐N) copolymers showed Tg over 130 °C as well as norbornene content over 50 mol %; both values were higher than those obtained by the cationic active species in 1 /[Ph3C][B(C6F5)4]. A series of copolymerization reactions by 1 /[Ph3C][B(C6F5)4]/AliBu3 without MWNTs produced copolymers with the same unexpected features. The NMR analysis revealed the presence of rac‐ENNE and rac‐ENNNE sequences. Thus, AliBu3 changed the stereoirregular alternating copolymer microstructure produced by 1 /[Ph3C][B(C6F5)4]. We conclude that AliBu3 is not only a scavenger for CNT impurities, but it reacts with the THF ligand to give coordinatively unsaturated active species. Finally, P(E‐co‐N)/MWNT masterbatches were mixed with commercial TOPAS to produce cyclic olefin copolymer nanocomposites with excellent dispersion of filler. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 5709–5719, 2009  相似文献   

6.
Synthesis and Properties of [Ph2(Carb)P]AlCl4 (Carb = 2,3‐Dihydro‐1,3‐diisopropyl‐4,5‐dimethylimidazol‐2‐ylidene) – a Stable Carbene Complex of Trivalent Phosphorus [1] 2,3‐Dihydro‐1,3‐diisopropyl‐4,5‐dimethylimidazol‐2‐ylidene ( 7 , Carb) reacts with chlorodiphenylphosphane to give the cationic phosphane [Ph2(Carb)P]Cl ( 10 ) which is transferred to the more stable salt [Ph2(Carb)]AlCl4 ( 13 ) on treatment with AlCl3. The cationic phosphane selenide [Ph2(Carb)PSe]AlCl4 ( 14 ) is obtained from 13 and selenium. Spectroscopic and structural data indicate [Ph2(Carb)P]+ to be a cationic analogue of Ph3P. The X‐ray structure of 13 is reported.  相似文献   

7.
Molybdenum disulfide (MoS2) is a promising candidate as a high‐performing anode material for sodium‐ion batteries (SIBs) due to its large interlayer spacing. However, it suffers from continued capacity fading. This problem could be overcome by hybridizing MoS2 with nanostructured carbon‐based materials, but it is quite challenging. Herein, we demonstrate a single‐step strategy for the preparation of MoS2 coupled with ordered mesoporous carbon nitride using a nanotemplating approach which involves the pyrolysis of phosphomolybdic acid hydrate (PMA), dithiooxamide (DTO) and 5‐amino‐1H‐tetrazole (5‐ATTZ) together in the porous channels of 3D mesoporous silica template. The sulfidation to MoS2, polymerization to carbon nitride (CN) and their hybridization occur simultaneously within a mesoporous silica template during a calcination process. The CN/MoS2 hybrid prepared by this unique approach is highly pure and exhibits good crystallinity as well as delivers excellent performance for SIBs with specific capacities of 605 and 431 mAhg?1 at current densities of 100 and 1000 mAg?1, respectively, for SIBs.  相似文献   

8.
The reaction of Ph2P(S)N(SiMe3)2 with potassium tert-butoxide in a 1:1 molar ratio produces K[Ph2P(S)NSiMe3], which was converted to the AsPh4+ salt by metathesis with [AsPh4]Cl. The X-ray crystal structure of [AsPh4][Ph2P(S)NSiMe3] · 0.5 THF consists of noninteracting AsPh4+ and Ph2P(S)NSiMe3? ions with d(P? S) = 1.980(4) Å and d(P? N) = 1.555(8) Å. The PNSi bite angle in the anion is 136.3(5)°. Electrophilic attack by Ph2P(S)Cl occurs at the sulfur atom of Ph2P(S)NSiMe3?. The oxidation of the anion with iodine produces a disulfide which regenerates K[Ph2P(S)NSiMe2] upon treatment with potassium tert-butoxide.  相似文献   

9.
Bunsen's cacodyl disulfide, Me2As(S)‐S‐AsMe2 ( 1 ), reacted with iodine giving the novel dimethylarsinosulfenyl iodide, Me2As‐S‐I ( 3 ) although theoretical calculations indicated that the AsV compound Me2As(S)‐I ( 4 ) was more stable in the gas phase. The oily product was stable neat and as a solution in CDCl3 at +4 °C and –20 °C for at least 15 d. Light, H2O, H2O2, and Zn dust, but not NaI or Ag, decomposed it. Compound 3 did not interact with Ph3N, with Ph2NH and PhNH2 it interacted but not reacted. 3 was decomposed by piperidine, with pyridine and 4‐dimethylaminopyridine it interacted and produced Me2As‐SS‐AsMe2 ( 2 ) and I2 that formed charge transfer complexes Base · I2, whereas Et3N decomposed 3 , and 3Et3N · 2I2 was isolated. 3 was desulfurized by Ph3P and (Me2N)3P completely, and by (PhO)3P and (PhS)3P partially. The reactions of 3 with (Me2N)3P, (PhS)3P, and (EtO)3P were complicated. From the AsIII nucleophiles, only Ph3As was bound, while (PhS)3As reacted slowly in a complicated manner with 3 . No interaction of 3 with MeOH or PhOH was observed but NaOH, Ag2O, and PhONa decomposed it. Thiophenol produced traces of Me2As‐SPh ( 10 ) and sodium thiophenolate attacked mainly at AsIII of 3 . Thus, externally stabilized sulfenium ions of the type Me2As‐S‐Nu+I were not obtained.  相似文献   

10.
A Contribution to Rhenium(II)‐, Osmium(II)‐, and Technetium(II)‐Thionitrosyl‐Complexes: Preparation, Structures, and EPR‐Spectra The reaction of [ReVINCl4] and [OsVINCl4] with S2Cl2 leads to the formation of the thionitrosyl complexes [MII(NS)Cl4] (M = Re, Os) which could not be isolated as pure compounds. Addition of pyridine to the reaction mixture results in the formation of the stable compounds trans‐(Ph4P)[OsII(NS)Cl4py], trans‐(Hpy)[OsII(NS)Cl4py], trans‐(Ph4P)[ReII(NS)Cl4py], and cis‐(Ph4P)[ReII(NS)Cl4py]. The crystal structure analyses show for trans‐(Ph4P)[OsII(NS)Cl4py] (monoclinic, P21/n, a = 12.430(3)Å, b = 18.320(4)Å, c = 15.000(3)Å, β = 114.20(3)°, Z = 4), trans‐(Hpy)[OsII(NS)Cl4py] (monoclinic, P21/n, a = 7.689(1)Å, b = 10.202(2)Å, c = 20.485(5)Å, β = 92.878(4)°, Z = 4), trans‐(Ph4P)[ReII(NS)Cl4py] (triclinic, P1¯, a = 9.331(5)Å, b = 12.068(5)Å, c = 15.411(5)Å, α = 105.25(1)°, β = 90.23(1)°, γ = 91.62(1)°, Z = 2), and cis‐(Ph4P)[ReII(NS)Cl4py] (monoclinic, P21/c, a = 10.361(1)Å, b = 16.091(2)Å, c = 17.835(2)Å, β = 90.524(2)°, Z = 4) M‐N‐S angles in the range 168‐175°. This indicates a nearly linear coordination of the NS ligand. The metal atom is octahedrally coordinated in all cases. The rhenium(II) thionitrosyl complexes (5d5 “low‐spin” configuration, S = 1/2) are studied by EPR in the temperature range 295 > T > 130 K. In addition to the detection of the complexes formed during the reaction of [ReVINCl4] with S2Cl2 EPR investigations on diamagnetically diluted powders and single crystals of the system (Ph4P)[ReII/OsII(NS)Cl4py] are reported. The 185, 187Re hyperfine parameters are used to get information about the spin‐density distribution of the unpaired electron in the complexes under study. [TcVINCl4] reacts with S2Cl2 under formation of [TcII(NS)Cl4] which is not stable and decomposes under S8 elimination and rebuilding of [TcVINCl4] as found by EPR monitoring of the reaction.  相似文献   

11.
The Influence of Phosphoryl Substituents on the Properties of P‐Substituted 2‐Methylimidazolium Ions and 2‐Methyleneimidazolines [1] The imidazolines ImCHP(E)Ph2 [ 6 , E = S ( a ), Se ( b )] are obtained from ImCHPPh2 ( 4 ) and sulfur or selenium. HBF4 reaction yields the corresponding imidazolium salts [ImCH2P(E)Ph2][BF4] [ 5 , E = S ( a ), Se ( b )]. 1, 3, 4, 5‐Tetramethyl‐2‐methylenimidazoline ( 1 , ImCH2) reacts with Ph2P(O)Cl to give the corresponding phosphane salt [ImCH2P(O)Ph2]Cl ( 7 ) from which the vinyl compound ImCHP(O)Ph2 ( 8 ) is formed through deprotonation. 8 reacts with excess HBF4 to give the phosphine oxide BF3 adduct [ImCH2P(O)Ph2 · BF3][BF4] ( 9 ). The crystal structures of 5a , 5b , 6b , 7 · CH2Cl2 and 9 · H2O as well as preliminary data of 8 are reported and discussed on comparison with the phosphanes [ImCH2PPh2][BF4] ( 3b ) and ImCHPPh2 ( 4 ). From structural data, π‐electron delocalisation is concluded for 6b and 8 .  相似文献   

12.
Treatment of 1,8‐bis(diphenylphosphino)naphthalene (dppn, 1 ) with stoichiometric amounts of sulfur or selenium in toluene at 80 °C selectively afforded the diphosphine monochalcogenides 1‐Ph2P(C10H6)‐8‐P(:S)Ph2 (dppnS, 2 a ) and 1‐Ph2P(C10H6)‐8‐P(:Se)Ph2 (dppnSe, 2 b ). The 31P{1H} NMR spectrum of 2 b showed an unusually large 5J(P–Se) value, which indicates a significant through‐space coupling component. The monosulfide acted as a bidentate P,S‐ligand towards platinum(II) ( 3 a ), whereas the corresponding monoselenide complex ( 3 b ′) lost elemental selenium with formation of the previously reported complex [PtCl2(dppn)‐P,P′] ( 3 ). Treatment of dppnSe with [(nor)Mo(CO)4] (nor = norbornadiene) led to formation of [(dppnSe)Mo(CO)4P,Se] ( 3 b ). Solutions of the latter slowly deposited Se with formation of [(dppn)Mo(CO)4P,P′] ( 4 ) which was also obtained by independent synthesis from 1 and [(nor)Mo(CO)4]. All isolated new compounds were characterised by a combination of 31P, 1H, 13C and 77Se ( 2 b ) NMR spectroscopy, IR spectroscopy, mass spectrometry and elemental analysis. Single‐crystal X‐ray structure determinations were performed for dppnSe ( 2 b ), [PtCl2(dppnS)‐P,S] ( 3 a ), [(dppnSe)Mo(CO)4P,Se] ( 3 b ) and [(dppn)Mo(CO)4P,P′] ( 4 ). In 2 b steric effects cause the naphthalene ring to be distorted and force the phosphorus atoms by 65 and 59 pm to opposite sides of the best naphthalene plane. In the metal complexes 3 a , 3 b and 4 the phosphino‐phosphinochalcogenyl systems act as bidentate ligands through the P and the chalcogen atoms. The naphthalene systems are again distorted. The two independent molecules of 4 differ in their conformations.  相似文献   

13.
Reactions of the oxorhenium(V) complexes [ReOX3(PPh3)2] (X = Cl, Br) with the N‐heterocyclic carbene (NHC) 1,3,4‐triphenyl‐1,2,4‐triazol‐5‐ylidene (LPh) under mild conditions and in the presence of MeOH or water give [ReOX2(Y)(PPh3)(LPh)] complexes (X = Cl, Br; Y = OMe, OH). Attempted reactions of the carbene precursor 5‐methoxy‐1,3,4‐triphenyl‐4,5‐dihydro‐1H‐1,2,4‐triazole ( 1 ) with [ReOCl3(PPh3)2] or [NBu4][ReOCl4] in boiling xylene resulted in protonation of the intermediately formed carbene and decomposition products such as [HLPh][ReOCl4(OPPh3)], [HLPh][ReOCl4(OH2)] or [HLPh][ReO4] were isolated. The neutral [ReOX2(Y)(PPh3)(HLPh)] complexes are purple, airstable solids. The bulky NHC ligands coordinate monodentate and in cis‐position to PPh3. The relatively long Re–C bond lengths of approximate 2.1Å indicate metal‐carbon single bonds.  相似文献   

14.
Inorganic graphene analogues (IGAs) are a huge and fascinating family of compounds that have extraordinary electronic, mechanical, and thermal properties. However, one of the largest problems that face the industrial application of IGAs is their poor processability, which has led to a “bottlenecking” in the development of freestanding, large‐area, IGA‐based thin‐film devices. Herein, we report a facile and cost‐efficient method to chemically modify IGAs by using their abundant coordination atoms (S, O, and N). Taking MoS2 as an example, we have prepared homogeneous “solution” systems, in which MoS2 nanosheets are chemically cross‐linked through a carboxylate‐containing polymeric ligand, poly(methyl methacrylate) (PMMA), by copper‐ion coordination. Bonding interactions between C?O bonds and sulfur atoms through copper ions were confirmed by various characterization techniques, such as UV/Vis, FTIR, and Raman spectroscopy and XPS. By using our method, freestanding MoS2 paper with significantly improved mechanical properties was obtained, thus laying the basis for the mass production of large‐area MoS2‐based thin‐film devices. Furthermore, copper‐ion coordination was also applied to MoS2/PMMA nanocomposites. Direct and strong nanofiller/matrix bonding interactions facilitate efficient load transfer and endow the polymeric nanocomposites with an excellent reinforcement effect. This method may pave a new way to high‐strength polymeric nanocomposites with superior frictional properties, flame retardance, and oxidation resistance.  相似文献   

15.
Fe? W heterometallic complexes, in which an FeX2 (X=Cl, SPh) moiety is attached to monodithiolene oxotungsten through a sulfide bridge, that is, [Ph4P]2[Cl2Fe(S)2WOS2] ( 1 ), [Ph4P]2[Cl2Fe(S)2WOS2(DMED)] ( 2 , DMED=dimethylethylenedicarboxylate), [Ph4P]2[Cl2Fe(S)2WO(tdt)] ( 3 , tdt=toluenedithiolate), [Ph4P]2[(SPh)2Fe(S)2WO(tdt)] ( 4 ), and [Ph4P]2[Cl2Fe(S)2WO(edt)] ( 5 , edt=ethanedithiolate), are reported. Mössbauer and EPR spectroscopy, magnetism, electrochemistry, and electronic structural analysis based on DFT and TD‐DFT calculations show the transfer of electron from the iron center to the tungsten center, thus resulting in a ferromagnetically coupled FeIIIWV unit, along with antiferromagnetic intermolecular interactions, from the starting FeII and WVI compounds. A net spin of a S=3 ground state, which arises from ferromagnetically coupled FeIII and WV atoms, displays a rare X‐band EPR in normal mode at g≈7 in the solid state.  相似文献   

16.
The alkaline earth metal complex [Mg{4,5‐(P(S)Ph2)2}2tz}2(thf)4] ( 2 ) and the bimetallic complexes, [M{4,5‐(P(S)Ph2)2}2tz}2(thf)]2 [M = Ca ( 3 ), Sr ( 4 ), Ba ( 5 )], [SrI{4,5‐(P(S)Ph2)2tz}2(thf)3]2 ( 6 ), and [{BaI(4,5‐(P(S)Ph2)2tz)}2(thf)7] ( 7 ) were prepared in good yields from the metathesis reactions of the potassium salt of 4,5‐bis(diphenylthiophosphoranyl)‐1,2,3‐triazole [H{4,5‐(P(S)Ph2)2tz}] ( 1 ) and MI2 (M = Ca, Sr, Ba), whereas the tetrametallic magnesium hydroxide [Mg2(μ‐OH){4,5‐(P(S)Ph2)2}2tz}3]2 ( 8 ) was obtained as the hydrolysis product from the starting material (MgnBuCl) and 1 . The NMR study of 2 – 8 in solution suggests the formation of solvated species in CD3OD‐d4, whereas for 4 , 5 , and 6 a fluxional behavior is observed in CD2Cl2. The structural analyses of 3 – 5 , 6 , and 7 in solid state reveal in all cases a central core defined by M2N4 heterocycle bearing M–S bonding. The degree of aggregation observed for these compounds depends significantly on the size of the metal atom as well as on the metal‐ligand molar ratio employed for each reaction.  相似文献   

17.
A series of aminodiphenylphosphanes 1 [Ph2P‐N(H)tBu ( a ), ‐NEt2 ( b ), ‐NiPr2 ( c )], 2 [Ph2P‐NHPh ( a ), ‐NH‐2‐pyridine ( b ), ‐NH‐3‐pyridine ( c ), ‐NH‐4‐pyridine ( d ), NH‐pyrimidine ( e ), NH‐2,6‐Me2‐C6H3 ( f ), NH‐3‐Me‐2‐pyridine ( g )], 3 [Ph2P‐N(Me)Ph ( a ), ‐NPh2 ( b )], and N‐pyrrolyldiphenylphosphane 4 (Ph2P‐NC4H4) was prepared and studied by NMR (1H, 13C, 31P, 15N NMR) spectroscopy. The isotope‐induced chemical shifts 1Δ14/15N(31P) were determined at natural abundance of 15N by using HEED INEPT experiments. A dependence of 1Δ14/15N(31P) on the substituents at nitrogen was found (alkyl < H < aryl; increasingly negative values). The magnitude and sign of the coupling constants 1J(31P,15N) (positive sign) are dominated by the presence of the lone pair of electrons at the phosphorus atom. The X‐ray structural analysis of 2b is reported, showing the presence of dimers owing to intermolecular hydrogen bridges in the solid state. © 2001 John Wiley & Sons, Inc. Heteroatom Chem 12:542–550, 2001  相似文献   

18.
Treatment of the thiosemicarbazone 2‐FC6H4C(Me)=NN(H)C(=S)NHPh, a , with palladium(II) acetate in acetic acid, or with lithium tetrachloropalladate(II) in methanol, gave the tetranuclear cyclometallated complex [Pd{2‐FC6H3C(Me)=NN=C(S)NHPh}]4 (1a) . Reaction of 1a with the diphosphines Ph2P(CH2)2PPh2 (dppe), Ph2PCH=CHPPh2 (trans‐dpe) Ph2P(CH2)3Ph2 (dppp) or Ph2P(CH2)4Ph2 (dppb) in a 1:2 molar ratio gave the dinuclear cyclometallated complexes [(Pd{2‐FC6H3C(Me)=NN=C(S)NHPh})2(μ‐Ph2P(CH2)nPPh2)], (n = 2, 2a ; 3, 4a ; 4, 5a ) and [(Pd{2‐FC6H3C(Me)=NN=C(S)NHPh})2(μ‐Ph2PCH=CHPPh2)], ( 3a ). The X‐ray crystal structure of ligand a and of complex 2a are described. The structure of complex 2a shows the palladium atom is bonded to four different donor atoms: C, N, S and P.  相似文献   

19.
In the two title complexes, (C24H20P)[Au(C3S5)2]·C3H6O, (I), and (C20H20P)[Au(C3S5)2], (II), the AuIII atoms exhibit square‐planar coordinations involving four S atoms from two 2‐thioxo‐1,3‐dithiole‐4,5‐dithiolate (dmit) ligands. The Au—S bond lengths, ranging from 2.3057 (8) to 2.3233 (7) Å in (I) and from 2.3119 (8) to 2.3291 (10) Å in (II), are slightly smaller than the sum of the single‐bond covalent radii. In (I), there are two halves of independent Ph4P+ cations, in which the two P atoms lie on twofold rotation axis sites. The Ph4P+ cations and [Au(C3S5)2] anions are interspersed as columns in the packing. Layers composed of Ph4P+ and [Au(C3S5)2] are separated by layers of acetone molecules. In (II), the [Au(C3S5)2] anions and EtPh3P+ counter‐cations form a layered arrangement, and the [Au(C3S5)2] anions form discrete pairs with a long intermolecular Au...S interaction for each Au atom in the crystal structure.  相似文献   

20.
Salts containing bis‐phosphonio‐benzophospholide cations 2 a – d with an additional donor site in one of the phosphonio‐moieties were synthesized either via quaternisation of the Ph2P moiety in the neutral phosphonio‐benzophospholide 3 , or via ring‐closure of the functionalized bis‐phosphonium ion 6 . The Ph2P‐substituted cation 2 d formed chelate complexes [M(k2P,P′‐ 2 d )(CO)n]+ with M(CO)n = Ni(CO)2, Fe(CO)3, Cr(CO)4. In the latter case, competition between formation of the chelate and a complex [Cr(kP‐ 2 d )2(CO)4]2+ was observed, and interpreted as a consequence of antagonism between the stabilizing chelate effect and destabilizing ligand–ligand repulsions. The formation of stable PdII and PtII complexes of 2 d suggests that the chelate effect may also overcome the kinetic inhibition which so far prevented isolation of complexes of these metals with bis‐phosphonio‐benzophospholides. The newly synthesized ligands and complexes were characterized by spectroscopic data, and an X‐ray crystal structure analysis of 2 a [Br]. The reactivity of chelate complexes towards Ph3P indicates that the ring phosphorus atom is a weaker donor than the pendant Ph2P‐group.  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