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 859 毫秒
1.
Internally consistent assignments of the 31P-{1H} NMR parameters of the complexes [Pt(RCCR′)(PPh3)2] are proposed, based on the premise that the magnitude of 1J(PtP) depends mainly on the nature of the moiety CR trans to P. For a given R, 2J(PP) correlates with 1J(PtP) for thebond trans to CR. The alkynes PhCCSnEt3, PhCCSnPh3, Me3SiCCCl, Me3SiCCBr, Et3SiCCI and MeCCI undergo oxidative addition reactions with [Pt(C2H4)(PPh3)2]; the intermediate alkyne complex was detected for PhCCSnEt3, Me3SiCCCl and Me3CCBr. The triyne Me(CC)3Me forms platinum(0) complexes by coordination with the central or terminal CC bond and appears also to give a platinum(II) complex by oxidative addition.  相似文献   

2.
Reaction of carbon diselenide in 3 to 1 molar ratio, and areneselenols in equimolar ratio, with trans-IrCl(CO)(PPPh3)2 and PtL4, gives oxidative addition products, IrCl(CO)CSe2)(PPh3)2, Pt(CSe2)L2, IrHCl(CO)(SeC6H4Me-p)(PPh3)2, and PtH(SeR)L2, respectively (R = Ph and p-MeC6H4; L = PPh3 and PPh2Me). However, reactions of PtL4 with an excess of areneselenols afford bis(arylselenide) complexes Pt(SeR)2L2. The configurations of these complexes are discussed on the basis of their IR and PMR spectra. The carbon diselenide adducts are suggested to have configurations similar to the corresponding carbon disulfide adducts. The platinum hydrides are found to exist as a mixture of cis and trans isomers in solution, both the isomers being labile with regard to dissociative exchange of the tertiary phosphine ligands. The trans configurations of Pt(SeR)2(PPh2Me)2 are unambiguously shown by the virtually coupled triplet pattern of the PPh2Me signals.  相似文献   

3.
Tetracloro-o-benzoquinone reacts with (diphenylacetylene)bis(tirphenylphosphine)platinum(0) to give the novel platinum(II) diphenylacetylene complex, Pt(C6Cl4O2)PhCCPh)(PPh3), (I), which reacts with hydrogen halides to give the compelexes cis-PtX2(PhCCPh((PPh3), (X = Cl or Br). Hydrogen chloride also readily removes the tetrachloro-o-benzoquinoneligand from the adducts Ni(C6Cl4O2)(Ph2PCH2CH2PPh2) and M(C6Cl4O2)(PPh3)2, (M = Pd or Pt) but it has no reaction upon Ir(Cl)(C6Cl4O2)(CO)(PPh3)2 at room temperature. The acetylene in (1) is susceptible to nucleophilic attact and reaction with diethylamine gives the vinyl adduct Pt(C6Cl4O2)(CPhCPh)NHEt2)(PPh3). Other reactions of (I) have also been studied. Attemps to prepare other olefin or acetylene complexes of platinum(II) by the action of tetrachlor-o-benzoquinone on the complexes Pt(L)(PPh3)2, (L = PhCCH,(Et)(Me)(HO)CCCC(OH)(Me)(Et), HOCH2OH, CF3CCCF3, CF2CF2, CF2CH2 or trans-PhCHCHPh) are also described.  相似文献   

4.
Treatment of Pt(PPh3)4 with N,N‐dimethylthiocarbamoyl chloride, Me2NC(=S)Cl, in dichloromethane at ?20 °C processes the oxidative addition reaction to produce platinum complex [Pt(PPh3)21‐SCNMe2)(Cl)], 2 with releasing two triphenylphosphine molecules. The 31P{1H} NMR spectra of complex 2 shows the dissociation of the triphenylphosphine ligand to form diplatinum complex [Pt(PPh3)Cl]2(μ,η2‐SCNMe2)2, 3 in which the two SCNMe2 ligands coordinated through carbon to one metal center and bridging the other metal through sulfur. Complex 2 is characterized by X‐ray diffraction analysis.  相似文献   

5.
Toxicity, antitumour, platinum distribution, hepatotoxicity and histology data are presented for a series of ferrocenylamines: [(η‐C5H4(CH2)nNH2)FeCp] (n = 0,1) ( 1 , 2 ); [(η‐C5H4CH2NHPh)FeCp] ( 3 ); [(η‐C5H4CH2NMe2)FeCp] ( 4 ); {[η‐C5H4CH(Me)NMe2]FeCp} ( 5 ); [η‐C5H4CH2NMe2)2Fe] ( 6 ); {[1,2η‐C5H3(CHMeNMe2)(PPh2)]FeCp} ( 7 ); {[1,2η‐C5H3(CHMeNMe2)(PPh2)]Fe[η‐C5H4PPh2]} ( 8 ); and their complexes cis‐PtCl2L2 ( 9 ); trans ‐ Pt(L)(dmso)X2 ( 10 ); [σ ‐ (L)Pt(dmso)X] ( 11 , 12 ) {σ‐(L)[Pt(dmso)X]2} ( 13 ); [σ‐(L)PtP(OPh)3Cl] ( 14 ) (L = ferrocenylamine). The toxicity order is 1 – 3 ≫ 4 – 8 for the ferrocenylamines; the lower toxicity of tertiary amines may be due to protonation in vivo. Pt(II) complexes all show increased toxicity over the ligand. Liver, not kidney, damage is the norm from i.p. injection of 1 – 14 and detailed platinum distribution, blood serum and histology studies with 9 and 11 show that the platinum distribution does not correlate with liver dysfunction. Complexes 9 – 14 , but not 1 – 8 , were active against P‐388 mouse leukaemia tumour and cisplatin‐resistant sarcoma, but inactive against L‐1210 mouse leukaemia and B‐16 melanoma. Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

6.
On the Reactivity of Alkylthio Bridged 44 CVE Triangular Platinum Clusters: Reactions with Bidentate Phosphine Ligands The 44 cve (cluster valence electrons) triangular platinum clusters [{Pt(PR3)}3(μ‐SMe)3]Cl (PR3 = PPh3, 2a ; P(4‐FC6H4)3, 2b ; P(n‐Bu)3, 2c ) were found to react with PPh2CH2PPh2 (dppm) in a degradation reaction yielding dinuclear platinum(I) complexes [{Pt(PR3)}2(μ‐SMe)(μ‐dppm)]Cl (PR3 = PPh3, 3a ; P(4‐FC6H4)3, 3b ; P(n‐Bu)3; 3e ) and the platinum(II) complex [Pt(SMe)2(dppm)] ( 4 ), whereas the addition of PPh2CH2CH2PPh2 (dppe) to cluster 2a afforded a mixture of degradation products, among others the complexes [Pt(dppe)2] and [Pt(dppe)2]Cl2. On the other hand, the treatment of cluster 2a with PPh2CH2CH2CH2PPh2 (dppp) ended up in the formation of the cationic complex [{Pt(dppp)}2(μ‐SMe)2]Cl2 ( 5 ). Furthermore, the terminal PPh3 ligands in complex 3a proved to be subject to substitution by the stronger donating monodentate phosphine ligands PMePh2 and PMe2Ph yielding the analogous complexes [{Pt(PR3)}2(μ‐SMe)(μ‐dppm)]Cl (PR3 = PMePh2, 3c ; PMe2Ph, 3d ). NMR investigations on complexes 3 showed an inverse correlation of Tolmans electronic parameter ν with the coupling constants 1J(Pt,P) and 1J(Pt,Pt). All compounds were fully characterized by means of NMR and IR spectroscopy. X‐ray diffraction analyses were performed for the complexes [{Pt{P(4‐FC6H4)3}}2(μ‐SMe)(μ‐dppm)]Cl ( 3b ), [Pt(SMe)2(dppm)] ( 4 ), and [{Pt(dppp)}2(μ‐SMe)2]Cl2 ( 5 ).  相似文献   

7.
The ligand 2,11-bis(diphenylphosphinomethyl)benzo[c]phenanthrene ( 1 ) has been used to prepare complexes of the type [PtL( 1 )] (L ? C2H4, CH2?CH? CO2Me, PhC?CPh, MeC?CMe, MeO2CC?CCO2Me, (i-Pr)O2CC?CCO2(i-Pr), Ph3P and CO). It is shown that these complexes are less labile than the corresponding species [PtL(Ph3P)2]. The preparation of complexes trans-[PtX(R)(1)] by oxidative addition of RX (RX ? PhCH2Br and Mel) to [Pt(C2H4)(1)] is described. The isolation of [PtO2(CH3)2CO(1)] is also reported.  相似文献   

8.
The coordination chemistry of platinum(II) with a series of thiosemicarbazones {R(H)C2=N3‐N2(H)‐C1(=S)‐N1H2, R = 2‐hydroxyphenyl, H2stsc; pyrrole, H2ptsc; phenyl, Hbtsc} is described. Reactions of trans‐PtCl2(PPh3)2 precursor with H2stsc (or H2ptsc) in 1 : 1 molar ratio in the presence of Et3N base yielded complexes, [Pt(η3‐ O, N3, S‐stsc)(PPh3)] ( 1 ) and [Pt(η3‐ N4, N3, S‐ptsc)(PPh3)] ( 2 ), respectively. Further, trans‐PtCl2(PPh3)2 and Hbtsc in 1 : 2 (M : L) molar ratio yielded a different compound, [Pt(η2‐ N3, S‐btsc)(η1‐S‐btsc)(PPh3)] ( 3 ). Complex 1 involved deprotonation of hydrazinic (‐N2H‐) and hydroxyl (‐OH) groups, and stsc2? is coordinating via O, N3, S donor atoms, while complex 2 involved deprotonation of hydrazinic (‐N2H‐) and ‐N4H groups and ptsc2? is probably coordinating via N4, N3, S donor atoms. Reaction of PdCl2(PPh3)2 with Hbtsc‐Me {C6H5(CH3)C2=N3‐N2(H)‐C1(=S)‐N1H2} yielded a cyclometallated complex [Pd(η3‐C, N3, S‐btsc‐Me)(PPh3)] ( 4 ). These complexes have been characterized with the help of analytical data, spectroscopic techniques {IR, NMR (1H, 31P), U.V} and single crystal X‐ray crystallography ( 1 , 3 and 4 ). The effects of substituents at C2 carbon of thiosemicarbazones on their dentacy and cyclometallation are emphasized.  相似文献   

9.
Yttrocene‐carboxylate complex [Cp*2Y(OOCArMe)] (Cp*=C5Me5, ArMe=C6H2Me3‐2,4,6) was synthesized as a spectroscopically versatile model system for investigating the reactivity of alkylaluminum hydrides towards rare‐earth‐metal carboxylates. Equimolar reactions with bis‐neosilylaluminum hydride and dimethylaluminum hydride gave adduct complexes of the general formula [Cp*2Y(μ‐OOCArMe)(μ‐H)AlR2] (R=CH2SiMe3, Me). The use of an excess of the respective aluminum hydride led to the formation of product mixtures, from which the yttrium‐aluminum‐hydride complex [{Cp*2Y(μ‐H)AlMe2(μ‐H)AlMe2(μ‐CH3)}2] could be isolated, which features a 12‐membered‐ring structure. The adduct complexes [Cp*2Y(μ‐OOCArMe)(μ‐H)AlR2] display identical 1J(Y,H) coupling constants of 24.5 Hz for the bridging hydrido ligands and similar 89Y NMR shifts of δ=?88.1 ppm (R=CH2SiMe3) and δ=?86.3 ppm (R=Me) in the 89Y DEPT45 NMR experiments.  相似文献   

10.
Displacement of norbornadiene (nbd; bicyclo[2.2.1]hepta‐2,5‐diene) from [Rh(PPh3)2(nbd)]ClO4 by hydrogenation in the presence of PPh3 and formamide or Me‐substituted derivatives, results in the formation of O‐bonded formamide complexes [Rh(PPh3)3(OCHNHxMe2−x)]ClO4 (x=0, 1, 2) rather than N‐bonded derivatives. These have been characterised by spectroscopic measurements and, in the case of [Rh(PPh3)3(OCHNHMe)]ClO4, by X‐ray crystallography. All undergo oxidative addition with H2, and the rates of ligand exchange in the RhI and RhIII complexes have been determined by magnetisation‐transfer measurements.  相似文献   

11.
Compounds of the general formula [ORe(OR)Cl2(PPh3)2] and [ORe(OEt)Cl2(PPh3)(py)], where R=alkyl or aryl and py=a substituted pyridine, were synthesized and their voltammetric behaviour investigated. For the former, the electron-transfer mechanism was observed to be dependent on solvent. In dry MeCN, a quasi-reversible oxidation and a reduction followed by a chemical reaction was observed. There were indications of nucleophilic attack on electrochemically generated [ORe(OEt)Cl2(PPh3)2]+, forming an unstable species whose reduction potentials were strongly dependent on the identity of the nucleophile. Voltammetric and spectroscopic observations of the oxorhenium(V) alkoxypyridine complex indicate the pyridine to be labile in halogenated hydrocarbon solvents but not in Me2CO, MeCN, or CCl4. Electrochemical generation of [ORe(OEt)Cl2(PPh3)(ClxCyHz)]+ (x=1,2, or 3; y=1 or 2; z=2,3, or 4) appears to be followed by transfer of a hydrogen atom from the solvent to form [(HO)Re(OEt)Cl2(PPh3)]+. Various pyridine complexes of this type were preparedvia substitution reactions under mild conditions. Varying the reaction conditions allowed the synthesis oftrans-dioxotetrapyridyl complexes in excellent yield.  相似文献   

12.
Transition Metal Silyl Complexes, 44. — Preparation of the Binuclear Silyl Complexes (CO)3(R3Si)Fe(μ-PR′R′′)Pt(PPh3)2 by Oxidative Addition of (CO)3(R′R′′HP)Fe(H)SiR3 to (C2H4)Pt(PPh3)2 The complexes (CO)3(R′R′′HP)Fe(H)SiR3 ( 1 ) [PHR′R′′ = PHPh2, PH2Ph, PH2Cy; SiR3 = SiPh3, SiPh2Me, SiPhMe2, Si(OMe)3] react with Pt(C2H4)(PPh3)2 to give the dinuclear, silyl-substituted complexes (CO)3(R3Si)Fe(μ-PR′R′′)Pt(PPh3)2 ( 2 ) in high yields. Upon reaction of 2 (R = R′ R′′ = Ph) with CO, the PPh3 ligand at Pt being trans to the PPh2 bridge is exchanged, and (CO)3(Ph3Si)Fe(μ-PPh2)Pt(PPh3)CO ( 3 ) is formed. Complex 3 is characterized by an X-ray structure analysis. The rather short Fe — Si distance [233.9(2) pm] and the infrared spectrum of 3 indicate that the Fe — Pt bond is quite polar.  相似文献   

13.
The work reports the unexpected reaction of diphenyldibromo antimonates (III) with PtCl2 and cis‐[PtCl2(PPh3)2]. The reaction gives triphenylstibine containing PtII complexes viz. cis‐[PtBr2(SbPh3)2] ( 1 ), trans‐[[PtBr(Ph)(SbPh3)2] ( 2 ), [NMe4][PtBr3(SbPh3)] ( 3 ), and cis‐[PtBr2(PPh3)(SbPh3)] ( 4 ). All the complexes were characterised by elemental analyses, IR, Raman, 195Pt NMR, FAB mass spectroscopy and X‐ray crystallography. A plausible mechanism via the phenyl migration is proposed for the formation of these complexes. The average Pt–Br distance in 1 is 2.456(2) Å, in 2 2.496 Å(trans to Ph) while in 3 it is 2.476 Å (trans to Sb) implying a comparable trans influence of Ph3Sb and Ph3P.  相似文献   

14.
195Pt, 119Sn and 31P NMR characteristics of the complexes trans-[Pt(SnCl3)(carbon ligand)(PEt3)2] (1a-1e) are reported, (carbon ligand = CH3 (1a), CH2Ph (1b), COPh (1c), C6Cl5 (1d), C6Cl4Y (e); Y = meta- and para-NO2, CF3, Br, H, CH3, OCH3, or Pt(SnCl3)(PEt3)2. The values of 1J(195Pt, 119Sn) vary from 2376 to 11895 Hz with the COPh ligand having the smallest and the C6Cl5 ligand the largest value, making a total range for this coupling constant, when the dimer syn-trans-[PtCl(SnCl3)(PEt3)]2 is included, of ca. 33000 Hz. In the meta- and para-substituted phenyl complexes 1J(195Pt, 119Sn) (a) is greater for electron-withdrawing substituents, (b) varies more for the meta-substituted derivatives (5634 to 7906 Hz) than for the para analogues (6088 to 7644 Hz) and (c) has the lowest values when the Pt(SnCl3)(PEt3)2 group is the meta- or para-substituent. The direction of the change in 1J(195Pt, 119Sn) is opposite to that found for 1J(195Pt, 119P). For the aryl complexes linear correlations are observed between δ(119Sn), 1J(195Pt, 119Sn), 1J(195Pt, 31P), 1J(119Sn, 31P) and the Hammett substituent constant σn. δ(119Sn) and 1J(195Pt, 119Sn) are related linearly to v(Pt-H) in the complexes trans-[PtH(C6H4Y)(PEt3)2]; δ(119Sn) and δ(1H) (hydride) are also linearly related. Based on 1J(195Pt, 119Sn), the acyl ligand is suggested to have a very large NMR trans influence. The differences in the NMR parameters for (1a-e) are rationalized in terms of differing σ- and π-bonding abilities of the carbon ligands.The structure of 1c has been determined by crystallographic methods. The complex has a slightly distorted square planar geometry with trans-PEt3 ligands. Relevant bond lengths (Å) and bond angles (°) are: PtSn, 2.634(1), PtP, 2.324(4) and 2.329(4), PtC, 2.05(1); PPtP, 170.7(6), SnPtC, 173.0(3), SnPtP, 92.1(1), 91.7(1), PPtC, 88.8(4) and 88.3(4). The PtSn bond separation is the longest yet observed for square-planar platinum trichlorostannate complexes, and would be consistent with a large crystallographic trans influence of the benzoyl ligand. The PtSn bond separation is shown to correlate with 1J(195Pt, 119Sn).  相似文献   

15.
Reaction of platinum(IV) chloride with SnCl2?·?2H2O in the presence of [NHR3]3Cl (R?=?Me, Et) in 3M hydrochloric acid affords the anionic five-coordinate platinum(II) complexes [NHR3]3[Pt(SnCl3)5], R?=?Me (1), Et (2), respectively. Moreover, platinum(IV) chloride reacts with SnCl2?·?2H2O in the presence of bis(triphenylphosphoranylidene)ammonium chloride in acetone/dichloromethane to form [N(PPh3)2]3[Pt(SnCl3)5] (3). In contrast, reaction of an acetone solution of platinum(IV) chloride with SnCl2?·?2H2O in the presence of bis(triphenylphosphoranylidene) ammonium chloride resulted in the formation of cis-[N(PPh3)2]2[PtCl2(SnCl3)2] (4). The same products are obtained by using a platinum(II) salt as starting material. Similarly, cis and trans- dichlorobis(diethyl sulfide)platinum(II) reacts with SnCl2?·?2H2O in 5M hydrochloric acid to give [PtCl(SEt2)3]3[Pt(SnCl3)5] (5) by facile insertion of SnCl2 into the Pt–Cl bond. However, treatment of an acetone solution of cis- and trans-[PtCl2(SEt2)2] with SnCl2?·?2H2O in the presence of a small amount of HCl resulted in the formation of 5, which dissociates in solution to give [PtCl2(SEt2)2]. The complexes have been fully characterized by elemental analysis and multinuclear NMR (1H,?13C,?195Pt,?119Sn) spectroscopy. A structure determination of crystals grown from a solution of 2 by X-ray diffraction methods shows that platinum adopts a regular trigonal bipyramidal geometry.  相似文献   

16.
Oxidative addition of 2‐phenylethylbromide (PhCH2CH2Br) to dimethylplatinum(II) complexes [PtMe2(NN)] ( 1a , NN = 2,2′‐bipyridine (bpy); 1b , NN = 1,10‐phenanthroline (phen)) afforded the new organoplatinum(IV) complexes [PtMe2(Br)(PhCH2CH2)(bpy)], as a mixture of trans ( 2a ) and cis ( 3a ) isomers, and [PtMe2(Br)(PhCH2CH2)(phen)], as a mixture of trans ( 2b ) and cis ( 3b ) isomers, respectively. The new Pt(IV) complexes were readily characterized using multinuclear (1H and 13C) NMR spectroscopy and elemental microanalysis. The crystal structure of 2a was further determined using X‐ray crystallography indicating an octahedral geometry around the platinum centre. A comparison of reactivity of RCH2Br reagents (R = CH3, Ph or PhCH2) in their oxidative addition reactions with complex 1a , with an emphasis on the effects of the R groups of alkyl halides, was also conducted using density functional theory.  相似文献   

17.
The complexes [Rh(X)(H)(SnPh3)(PPh3)(L)] (X = NCBPh3 (a), N(CN)2 (b), NCS (c), NCO (d), N3 (e); L = 1‐methylimidazole) ( 1 ) show systematic changes in δ(119Sn), δ(103Rh), J(119Sn–1H) and J(119Sn–103Rh) that are related to the electron‐donating properties of X. As X becomes more electron‐rich, δ(103Rh), J(119Sn–1H) and J(119Sn–103Rh) increase and δ119Sn) decreases. The related complexes trans‐[Rh(X)(H)(SnPh3)(PPh3)2(L)] (X = N(CN)2, NCO; L = 4‐carboxymethylpyridine (x), pyridine (y) and 4‐dimethylaminopyridine (z)) ( 2 ), show a continuation of the trends in δ(119Sn) and J(119Sn–1H), but not δ(103Rh) or J(119Sn–103Rh). Data for 1 and 2 show that within certain limits of type of ligand varied (X = N‐donor, L = a pyridine) and coordination geometry, the response of δ(119Sn) and J(119Sn–1H) to changes in electron density on rhodium is largely independent of the means by which the change is effected.Copyright © 2001 John Wiley & Sons, Ltd.  相似文献   

18.
5-C5Me5)(CO)2(PPh3)MoCHO (2) one of the few isolated neutral metal formyls, reacts with the electrophilic reagents (CF3COOH and CH3SO3F without disproportionation to give the secondary carbene complexes [(η5-C5Me5)(CO)2(PPh3)Mo(CHOE)]+ X (E = H, X = CF3COO (4); E = Me, X = PF6 (5)).  相似文献   

19.
The unsaturated complexes RuCl(CO)(RC=CHR′)(PPh3)2 react with CO to give the dicarbonyl complexes RuCl(CO)2(RC=CHR′)(PPh3)2 or the η2-acyl complexes RuCl(CO)(O=CC(R)=CHR′)(PPh3)2, depending on the R and R′ groups. The RuCl(CO)(O=CC(Me)=CHMe)(PPh3)2 complex reacts with methanol to give RuCl(CO)(O2CC(Me)=CHMe)(PPh3)2, which structure has been established by an X-ray diffraction study.  相似文献   

20.
The redox reaction of bis(2-benzamidophenyl) disulfide (H2L-LH2) with [Pd(PPh3)4] in a 1:1 ratio gave mononuclear and dinuclear palladium(II) complexes with 2-benzamidobenzenethiolate (H2L), [Pd(H2L-S)2(PPh3)2] (1) and [Pd2(H2L-S)2 (μ-H2L-S)2(PPh3)2] (2). A similar reaction with [Pt(PPh3)4] produced only the corresponding mononuclear platinum(II) complex, [Pt(H2L-S)2(PPh3)2] (3). Treatment of these complexes with KOH led to the formation of cyclometallated palladium(II) and platinum(II) complexes, [Pd(L-C,N,S)(PPh3)] ([4]) and [Pt(L-C,N,S) (PPh3)] ([5]). The molecular structures of 2, 3 and [4] were determined by X-ray crystallography.  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