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.
《Polyhedron》2004,23(2-3):273-282
Sodium salt of the fluorinated tris(pyrazolyl)borate ligand [HB(3-(CF3),5-(Ph)Pz)3] has been synthesized from the corresponding pyrazole and NaBH4 in high yield. It forms stable adducts with oxygen-based donors like THF and water. [HB(3-(CF3),5-(Ph)Pz)3]Na(THF) is monomeric in the solid state whereas {[HB(3-(CF3),5-(Ph)Pz)3]Na(H2O)}n forms a polymeric chain structure. These tris(pyrazolyl)boratosodium adducts show intra and/or inter-molecular sodium–fluorine interactions. X-ray crystal structures of [HB(3,5-(CF3)2Pz)3]Na(OEt2) and {[HB(3-(CF3),5-(CH3)Pz)3]Na(H2O)}2 also display similar CF⋯Na contacts. The copper carbon monoxide complex [HB(3-(CF3),5-(Ph)Pz)3]CuCO was prepared by treating [HB(3-(CF3),5-(Ph)Pz)3]Na(THF) with CuOTf in the presence of CO. The infra-red stretching frequency data show relatively high νCO value (2103 cm−1) indicating the presence of highly electrophilic copper site on [HB(3-(CF3),5-(Ph)Pz)3]CuCO. Crystal structure of 3-(CF3),5-(Ph)PzH is also reported. It forms “tub” shaped tetramers in the solid state.  相似文献   

2.
The interactions of cyclic trinuclear copper {[3,5-(CF3)2Pz]Cu}3 and silver {[3,5-(CF3)2Pz]Ag}3 complexes with polyhedral borate anions [B10H10]2− and [B12H12]2− in solvents of low-polarity were studied using IR spectroscopy (190-290 K). Two types of complexes were found in solution: {[((3,5-CF3)2PzM)3][BnHn]}2− and {[((3,5-CF3)2PzM)3]2[BnHn]}2− (M = Ag, Cu; n = 10, 12). The stability constants of these complexes were determined from IR-spectra.  相似文献   

3.
Reactions of the sterically bulky mono-valent group 13 bisimidinate gallium(I), Ga(DDP) (1) (DDP = 2-{(2, 6-diisopropylphenyl)amino}-4-{(2, 6-diisopropylphenyl)imino}-2-pentene, HC(CMeNC6H3-2,6-iPr2)2) with olefin supported group 10 complexes, [(diene)PtCl2] [diene = 1,5-cyclooctadiene (COD), endo-dicyclopentadiene (dcy)] and [(COD)Pd(Me)(OTf)] (OTf = O3SCF3) are reported. These reactions afforded [(COD)Pt(Cl){ClGa(DDP)}] (2), [(dcy)Pt(Cl){ClGa(DDP)}] (3) and [(DDP)Ga(Me)(OTf)] (4) in moderate yields. Compounds 2-4 were characterized by elemental analysis, NMR (1H, 13C) spectroscopy and also by single crystal X-ray structural analysis. The solid state structures of complexes 2 and 3 reveal the oxidative insertion of Ga(DDP) into the Pt-Cl bond without altering the π-coordinated double bonds in the olefin.  相似文献   

4.
The dipalladium complexes, [PdCl(μ-MeN{P(OR)2}2)]2 (R = CH2CF3, 1a; Ph, 1b) react with [Mo25-C5H5)2(CO)6] in boiling benzene to afford the molybdenum-palladium heterometallic complexes, [(η5-C5H5)(CO)Mo(μ-MeN{P(OR)2}2)2PdCl] (R = CH2CF3, 3a; Ph, 3b), [(η5-C5H5)Mo(μ3-CO)2(μ-MeN{P(OR)2}2)2Pd2Cl], (R = CH2CF3, 5a; Ph, 5b), [(η5-C5H5)(Cl)Mo(μ2-CO)(μ2-Cl)(μ-MeN{P(OR)2}2)PdCl], (R = CH2CF3, 6a; Ph, 6b) and also the mononuclear complex [Mo(CO)Cl(η5-C5H5)(κ2-MeN{P(OR)2}2)], (R = Ph, 4b). These complexes have been separated by column chromatography and are characterised by elemental analysis, IR, 1H, 31P{1H} NMR data. The structures of 1a, 3a, 4b, 5b and 6a have been confirmed by single crystal X-ray diffraction. The CO ligands in 5b and 6a adopt a semi-bridging mode of bonding; the Mo-CO distances (1.95-1.97 Å) are shorter than the Pd-CO distances (2.40-2.48 Å). The Pd-Mo distances fall in the range, 2.63-2.86 Å. The reaction of [Mo25-C5H5)2(CO)6] with MeN{P(OPh)2}2 in toluene gives [Mo2(CO)45-C5H5)21-MeN{P(OPh)2}2)2] (2) in which the diphosphazane acts as a monodentate ligand.  相似文献   

5.
A library of rare-earth metal derivatives supported by an aminophenoxy ligand was prepared and their catalytic performance in lactide polymerization was investigated. It was found that the synthetic strategy had a profound effect on the formation of aminophenoxy rare-earth metal complexes. Amine elimination between Ln[N(SiMe3)2]3(μ-Cl)Li(THF)3 (Ln = Yb, Y) and 1 equiv. of the aminophenol [HONH] ([HONH] = ο-OCH3-C6H4NHCH2(3,5-tBu2-C6H2-2-OH)) in toluene gave the unexpected heterobimetallic bis(aminophenoxy) rare-earth metal complexes [ON]2LnLi(THF)2 (Ln = Yb ( 1 ), Y ( 2 )). When the reactions were carried out in THF and TMEDA, amine elimination produced the aminophenoxy rare-earth metal amide complexes {[ON]LnN(SiMe3)2}2 (Ln = Yb ( 5 ), Y ( 6 )) in ca 85% isolated yields. Complexes 5 and 6 could also be obtained from salt metathesis reaction of {[ON]LnCl(THF)}2 (Ln = Yb ( 3 ), Y ( 4 )) with NaN(SiMe3)2 in a 1:2 molar ratio. In addition, treatment of complexes 3 and 4 with NaOAr (Ar = &bond;C6H4-4-tBu) and (SiMe3)2NC(NPri)2Na in 1:4 and 1:2 molar ratios provided the corresponding aminophenoxy rare-earth metal derivatives {[ON](μ-OAr)Ln(μ-OAr)Na(THF)2}2 (Ln = Yb ( 7 ), Y ( 8 )) and {[ON]Ln[(iPrN)2CN(SiMe3)2]}2 (Ln = Yb ( 9 ), Y ( 10 )), respectively. These complexes were fully characterized, and their molecular structures were determined using single-crystal X-ray diffraction. Polymerization experiments showed that complexes 1 , 2 , 5 , 6 , 9 and 10 were highly active for the ring-opening polymerization of l -lactide in toluene, and complex 1 promoted l -lactide polymerization in a controlled fashion. The polymerization of rac-lactide initiated by the neutral aminophenoxy rare-earth metal complexes 5 , 6 , 9 and 10 in THF afforded heterotactic polymers.  相似文献   

6.
The reaction of the lithium salt of backbone fluorinated β‐diketiminate ligands, ArNC(CF3)CHC(CF3) NArLi, with trans‐[NiCl(Ph)(PPh3)2] gives nickel (II) complexes, ArNC(CF3)CHC(CF3)NAr(Ph) (PPh3)Ni (Ar = 2, 6‐Me2C6H3: 1 ; 2, 6‐iPr2C6H3: 2 ). When activated by methylaluminoxane (MAO), both complexes polymerize norbornene rapidly via a vinyl‐type polymerization mechanism. Treatment of nickel complex 1 with oxygen gives rise to intramolecular aerobic hydroxylation. The oxygenated species 3 was characterized by X‐ray crystallography. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

7.
The pincer complex [(POCOP)Ni(NCMe)][OSO2CF3] (1: POCOP = {2,6-(i-Pr2PO)2C6H3}) undergoes an acetonitrile substitution reaction in the presence of CN(t-Bu), KCN, and KOCN to give the new complexes [(POCOP)Ni{CN(t-Bu)}][O3SCF3] and (POCOP)Ni(X) (X = CN and NCO). The Ni-CN derivative is also obtained from a gradual decomposition of the Ni-CN(t-Bu) derivative, while the aquo derivative [(POCOP)Ni(OH2)][O3SCF3] was obtained from slow hydrolysis of (POCOP)Ni(OSO2CF3). All new complexes have been characterized spectroscopically and by X-ray crystallography. IR and solid state structural data indicate that Ni-L/X interactions are dominated by ligand-to-metal σ-donation; presence of little or no π-backbonding is consistent with the electrophilicity of the cationic fragment [(POCOP)Ni]+.  相似文献   

8.
Enantiomerically pure triflones R1CH(R2)SO2CF3 have been synthesized starting from the corresponding chiral alcohols via thiols and trifluoromethylsulfanes. Key steps of the syntheses of the sulfanes are the photochemical trifluoromethylation of the thiols with CF3Hal (Hal=halide) or substitution of alkoxyphosphinediamines with CF3SSCF3. The deprotonation of RCH(Me)SO2CF3 (R=CH2Ph, iHex) with nBuLi with the formation of salts [RC(Me)? SO2CF3]Li and their electrophilic capture both occurred with high enantioselectivities. Displacement of the SO2CF3 group of (S)‐MeOCH2C(Me)(CH2Ph)SO2CF3 (95 % ee) by an ethyl group through the reaction with AlEt3 gave alkane MeOCH2C(Me)(CH2Ph)Et of 96 % ee. Racemization of salts [R1C(R2)SO2CF3]Li follows first‐order kinetics and is mainly an enthalpic process with small negative activation entropy as revealed by polarimetry and dynamic NMR (DNMR) spectroscopy. This is in accordance with a Cα? S bond rotation as the rate‐determining step. Lithium α‐(S)‐trifluoromethyl‐ and α‐(S)‐nonafluorobutylsulfonyl carbanion salts have a much higher racemization barrier than the corresponding α‐(S)‐tert‐butylsulfonyl carbanion salts. Whereas [PhCH2C(Me)SO2tBu]Li/DMPU (DMPU = dimethylpropylurea) has a half‐life of racemization at ?105 °C of 2.4 h, that of [PhCH2C(Me)SO2CF3]Li at ?78 °C is 30 d. DNMR spectroscopy of amides (PhCH2)2NSO2CF3 and (PhCH2)N(Ph)SO2CF3 gave N? S rotational barriers that seem to be distinctly higher than those of nonfluorinated sulfonamides. NMR spectroscopy of [PhCH2C(Ph)SO2R]M (M=Li, K, NBu4; R=CF3, tBu) shows for both salts a confinement of the negative charge mainly to the Cα atom and a significant benzylic stabilization that is weaker in the trifluoromethylsulfonyl carbanion. According to crystal structure analyses, the carbanions of salts {[PhCH2C(Ph)SO2CF3]Li? L }2 ( L =2 THF, tetramethylethylenediamine (TMEDA)) and [PhCH2C(Ph)SO2CF3]NBu4 have the typical chiral Cα? S conformation of α‐sulfonyl carbanions, planar Cα atoms, and short Cα? S bonds. Ab initio calculations of [MeC(Ph)SO2tBu]? and [MeC(Ph)SO2CF3]? showed for the fluorinated carbanion stronger nC→σ* and nO→σ* interactions and a weaker benzylic stabilization. According to natural bond orbital (NBO) calculations of [R1C(R2)SO2R]? (R=tBu, CF3) the nC→σ*S? R interaction is much stronger for R=CF3. Ab initio calculations gave for [MeC(Ph)SO2tBu]Li ? 2 Me2O an O,Li,Cα contact ion pair (CIP) and for [MeC(Ph)SO2CF3]Li ? 2 Me2O an O,Li,O CIP. According to cryoscopy, [PhCH2C(Ph)SO2CF3]Li, [iHexC(Me)SO2CF3]Li, and [PhCH2C(Ph)SO2CF3]NBu4 predominantly form monomers in tetrahydrofuran (THF) at ?108 °C. The NMR spectroscopic data of salts [R1(R2)SO2R3]Li (R3=tBu, CF3) indicate that the dominating monomeric CIPs are devoid of Cα? Li bonds.  相似文献   

9.
Synthesis, Structure, and Reactivity of the Ferrioarsaalkene [(η5‐C5Me5)(CO)2FeAs=C(Ph)NMe2] Reaction of equimolar amounts of the carbenium iodide [Me2N(Ph)CSMe]I and LiAs(SiMe3)2 · 1.5 THF afforded the thermolabile arsaalkene Me3SiAs = C(Ph)NMe2 ( 1 ), which in situ was converted into the black crystalline ferrioarsaalkene [(η5‐C5Me5)(CO)2FeAs=C(Ph)NMe2)] ( 2 ) by treatment with [(η5‐C5Me5)(CO)2FeCl]. Compound 2 was protonated by ethereal HBF4 to yield [(η5‐C5Me5)(CO)2FeAs(H)C(Ph)NMe2]BF4 ( 3 ) and methylated by CF3SO3Me to give [(η5‐C5Me5)(CO)2FeAs(Me)C(Ph)NMe2]‐ SO3CF3 ( 4 ). [(η5‐C5Me5)(CO)2FeAs[M(CO)n]C(Ph)NMe2] ( 5 : [M(CO)n] = [Fe(CO)4]; 6 : [Cr(CO)5]) were isolated from the reaction of 2 with [Fe2(CO)9] or [{(Z)‐cyclooctene}Cr(CO)5], respectively. Compounds 2 – 6 were characterized by means of elemental analyses and spectroscopy (IR, 1H, 13C{1H}‐NMR). The molecular structure of 2 was determined by X‐ray diffraction analysis.  相似文献   

10.
The reactions of PhCboSeNa (Cbo = o-C2B10H10), prepared by reductive cleavage of Se-Se bond in (PhCboSe)2 by NaBH4 in methanol, with Na2PdCl4, MCl2(PR3)2 and [M2Cl2(μ-Cl)2(PR3)2] afforded a variety of complexes, viz., [Pd(SeCboPh)Cl] (1), [M(SeCboPh)2(PR3)2], [M2Cl2(μ-SeCboPh)(μ-Cl)(PR3)2] (M = Pd, Pt) and [Pd2Cl(SeCb0Ph)(μ-Cl)(μ-SeCboPh)(PEt3)2] (7) have been isolated. These complexes were characterized by elemental analyses and NMR (1H, 31P, 77Se, 195Pt) spectroscopy. The structures of [Pd(SeCboPh)2(PEt3)2] (2), [Pt(SeCboPh)2(PMe2Ph)2] (3), [Pd2Cl2(μ-SeCboPh)(μ-Cl)(PMe2Ph)2] (5) and [Pd2Cl(SeCboPh)(μ-Cl)(μ-SeCboPh)(PEt3)2] (7) were established by X-ray crystallography. The latter represents the first example of asymmetric coordination of selenolate ligands in binuclear bis chalcogenolate complexes of palladium and platinum. Thermolysis of [Pd(SeCboPh)2(PEt3)2] (2) in HDA (hexadecylamine) at 330 °C gave nano-crystals of Pd17Se15.  相似文献   

11.
The reaction of CuI, AgI, and AuI salts with carbon monoxide in the presence of weakly coordinating anions led to known and structurally unknown non‐classical coinage metal carbonyl complexes [M(CO)n][A] (A=fluorinated alkoxy aluminates). The coinage metal carbonyl complexes [Cu(CO)n(CH2Cl2)m]+[A]? (n=1, 3; m=4?n), [Au2(CO)2Cl]+[A]?, [(OC)nM(A)] (M=Cu: n=2; Ag: n=1, 2) as well as [(OC)3Cu???ClAl(ORF)3] and [(OC)Au???ClAl(ORF)3] were analyzed with X‐ray diffraction and partially IR and Raman spectroscopy. In addition to these structures, crystallographic and spectroscopic evidence for the existence of the tetracarbonyl complex [Cu(CO)4]+[Al(ORF)4]? (RF=C(CF3)3) is presented; its formation was analyzed with the help of theoretical investigations and Born–Fajans–Haber cycles. We discuss the limits of structure determinations by routine X‐ray diffraction methods with respect to the C? O bond lengths and apply the experimental CO stretching frequencies for the prediction of bond lengths within the carbonyl ligand based on a correlation with calculated data. Moreover, we provide a simple explanation for the reported, partly confusing and scattered CO stretching frequencies of [CuI(CO)n] units.  相似文献   

12.
Two CoII complexes, namely {[CoL(MeOH)(μ-OAc)]2Co}·2MeCN·2MeOH (1) and {[CoL(EtOH)(μ-OAc)]2Co}·3EtOH (2) (H2L=3,3′-dimethoxy-2,2′-[(1,3-propylene)dioxybis(nitrilomethylidyne)]diphenol), have been synthesized and characterized by X-ray crystallography. Both complexes contain octahedral coordination geometries, comprising three CoII atoms, two deprotonated bisoxime L2− units in which four μ-phenoxo oxygen atoms form two [CoL(X)] (X = MeOH or EtOH) units, two acetate ligands coordinated to three CoII centers through Co–O–C–O–Co bridges, and coordinated and non-coordinated solvent. Both complexes exhibit 2D supramolecular networks through different intermolecular hydrogen-bonding interactions.  相似文献   

13.
Four pyridinecarboxamide iron dicyanide building blocks and one Mn(III) compound have been employed to assemble cyanide-bridged heterometallic complexes, resulting in a series of trinuclear cyanide-bridged FeIII–MnII complexes: {[Mn(DMF)2 (MeOH)2][Fe(bpb)(CN)2]2}·2DMF (1), {[Mn(MeOH)4][Fe(bpmb)(CN)2]2}·2MeOH·2H2O (2), {[Mn(MeOH)4][Fe(bpdmb)(CN)2]2}·2MeOH·2H2O (3) and {[Mn(MeOH)4][Fe(bpClb)(CN)2]2}·4MeOH (4) (bpb2− = 1,2-bis(pyridine-2-carboxamido)benzenate, bpmb2− = 1,2-bis(pyridine-2-carboxamido)-4-methyl-benzenate, bpdmb2− = 1,2-bis(pyridine-2-carboxamido)-4,5-dimethyl-benzenate, bpClb2− = 1,2-bis(pyridine-2-carboxamido)-4-chloro-benzenate). Single-crystal X-ray diffraction analysis shows their similar sandwich-like structures, in which the two cyanide-containing building blocks act as monodentate ligands through one of their two cyanide groups to coordinate the Mn(II) center. Investigation of the magnetic properties of these complexes reveals antiferromagnetic coupling between the neighboring Fe(III) and Mn(II) centers through the bridging cyanide group. A best fit to the magnetic susceptibilities of complexes 1 and 3 gave the magnetic coupling constants J = −1.59(2) and −1.32(4) cm−1, respectively.  相似文献   

14.
The gold sulfonium benzylide complexes [( P1 )AuCHPh(SR1R2)]+ {B[3,5-CF3C6H3]4} [ P1 =P(tBu)2o-biphenyl; R1, R2=-(CH2)4- ( 1 a ); R1=Et, R2=Ph ( 1 b ); R1=R2=Ph ( 1 c )] were synthesized by reaction of the gold α-chloro benzyl complex ( P1 )AuCHClPh with sodium tetrakis[3,5-bis(trifluoromethyl)phenyl]borate and excess sulfide. Complexes 1 undergo efficient benzylidene transfer to alkenes and DMSO under mild conditions without external activation. Kinetic analysis of the reaction of 1 c with styrene was consistent with the intermediacy of the cationic gold benzylidene complex [( P1 )AuCHPh]+ ( I ).  相似文献   

15.
The reactions of [MCl2(PP)] and [MCl2(PR3)2)] with 1-mercapto-2-phenyl-o-carborane/NaSeCboPh and 1,2-dimercapto-o-carborane yield mononuclear complexes of composition, [M(SCboPh)2(PP)], [M(SeCboPh)2(PP)] (M = Pd or Pt; PP = dppm (bis(diphenylphosphino)methane), dppe (1,2-bis(diphenylphosphino)ethane) or dppp (1,3-bis(diphenylphosphino)propane)) and [M(SCboS)(PR3)2] (2PR3 = dppm, dppe, 2PEt3, 2PMe2Ph, 2PMePh2 or 2PPh3). These complexes have been characterized by elemental analysis and NMR (1H, 31P, 77Se and 195Pt) spectroscopy. The 1J(Pt–P) values and 195Pt NMR chemical shifts are influenced by the nature of phosphine as well as thiolate ligand. Molecular structures of [Pt(SCboPh)2(dppm)], [Pt(SeCboPh)2(dppm)], [Pt(SCboS)(PMe2Ph)2] and [Pt(SCboS)(PMePh2)2] have been established by single crystal X-ray structural analyses. The platinum atom in all these complexes acquires a distorted square planar configuration defined by two cis-bound phosphine ligands and two chalcogenolate groups. The carborane rings are mutually anti in [Pt(SCboPh)2(dppm)] and [Pt(SeCboPh)2(dppm)].  相似文献   

16.
Chloride abstraction from [(R,R)‐(iPrDuPhos)Co(μ‐Cl)]2 with NaBArF4 (BArF4=B[(3,5‐(CF3)2)C6H3]4) in the presence of dienes, such as 1,5‐cyclooctadiene (COD) or norbornadiene (NBD), yielded long sought‐after cationic bis(phosphine) cobalt complexes, [(R,R)‐(iPrDuPhos)Co(η22‐diene)][BArF4]. The COD complex proved substitutionally labile undergoing diene substitution with tetrahydrofuran, NBD, or arenes. The resulting 18‐electron, cationic cobalt(I) arene complexes, as well as the [(R,R)‐(iPrDuPhos)Co(diene)][BArF4] derivatives, proved to be highly active and enantioselective precatalysts for asymmetric alkene hydrogenation. A cobalt–substrate complex, [(R,R)‐(iPrDuPhos)Co(MAA)][BArF4] (MAA=methyl 2‐acetamidoacrylate) was crystallographically characterized as the opposite diastereomer to that expected for productive hydrogenation demonstrating a Curtin–Hammett kinetic regime similar to rhodium catalysis.  相似文献   

17.
By using the macrocyclic oxamido-copper complex CuL (H2L = 2,3-dioxo-5,6:13,14-dibenzo-9,10-(O)cyclohexyl-1,4,8,11-tetraazacyclo-tetradeca-7,12-diene) as precursor, two new trinuclear complexes with the formulas [(CuL)2Mn(ClO4)2] (1) and [(CuL)2Co(ClO4)2] (2) have been synthesized and structurally characterized. H-bonds are found between the molecules, which link adjacent trinuclear units together to form a unique one-dimensional structure. The temperature dependence of the magnetic susceptibility for the complexes was analyzed by means of the Hamiltonian leading to J = −14.66 cm−1 and J = −22.9 cm−1 for 1 and 2, respectively. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

18.
Two new optically active bidentate N,N‐ligands, DMIQCI ( 3a ) and DMIQCD ( 3b ), containing a quinuclidine core and an imidazolidin‐2‐imine unit, were synthesized. The reaction of these ligands with [(η5‐C5Me5)RuCl]4 afforded the brick‐red ruthenium(II) complexes [(η5‐C5Me5)Ru(DMIQCI)Cl] ( 4 ) and [(η5‐C5Me5)Ru(DMIQCD)Cl] ( 5 ), which were used as catalysts in the transfer hydrogenation of acetophenone in boiling 2‐propanol. The reactions of 3a and 3b with [(COD)PdCl2] (COD = 1,5‐cycloocta‐diene) and with [(DME)NiBr2] (DME = 1,2‐dimethoxyethane) afforded the square‐planar palladium(II) complexes [(DMIQCI)PdCl2] ( 7 ) and [(DMIQCD)PdCl2] ( 8 ) or the tetrahedral nickel(II) complexes [(DMIQCI)NiBr2] ( 9 ) and [(DMIQCD)NiBr2] ( 10 ), respectively. The X‐ray crystal structures of 4 , 7 , 9· THF, and 10 are reported.  相似文献   

19.
The CH bonds of tetrahydrofuran, acetone and benzene are activated by the fragment [(PP3)Ir]+ (PP3 = P(CH2CH2PPh2)3) generated either by photolytic dehydrogenation of the cis-dihydride [(PP3)IrH2](SO3CF3) or by thermal decomposition of the cis-organyl)hydrides [(PP)3)IrH(R)](SO3CF3) (R = Me, Ph). The latter compounds are obtained by protonation of the σ-organyl complexes (PP3)IrR.  相似文献   

20.
The catalytic system Pd(CF3COO)2-Ph2P(CH2)3PPh2-MeOH/Me2CO was studied by electrospray mass spectrometry. The {[Pd(dPPP)2]2+ [(dppp)2Pd(CF3COO)]+, [(dppp)Pd(CF3COO)]+, and [(dppp)Pd(CF3COO)2Pd(dppp)]2+} cations were found in the system. The addition of H2O to the system resulted in the formation of binuclear bicharged ions [(dppp)Pd(OH)]2 2+ and their associates with water.Translated from Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 2, pp. 494–496, February, 1996  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