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.
Photolysis of [Ir(η2-coe)H2(TpMe2)] ( 1 ; TpMe2=hydrotris(3,5-dimethylpyrazolyl)borato, coe=(Z)-cyclooctene) in CH3OH gives a mixture of [IrH4(TpMe2)] ( 4 ) and [Ir(CO)H2(TpMe2)] ( 5 ) in a ca. 1 : 1 ratio. Mass-spectral analysis of the distillate of the reaction mixture at the end of the photolysis shows the presence of coe. When pure CD3OD is used as solvent, the deuteride complexes [IrD4(TpMe2)] ((D4)- 4 ) and [Ir(CO)D2(TpMe2)] ((D2)- 5 ) are obtained. Also the photolysis of [Ir(η4-cod)(TpMe2)] ( 3 ) (cod=cycloocta-1,5-diene) gives 4 and 5 . A key feature of this photoreaction is the intramolecular dehydrogenation of cod with formation of cycloocta-1,3,5-triene, detected by mass spectroscopy at the end of the photolysis. Labeling experiments using CD3OD show that the hydrides in 4 originate from MeOH. When 13CH3OH is used as solvent, [Ir(13CO)H2(TpMe2)] is formed demonstrating that CH3OH is the source of the CO ligand. The observation that the photolysis of both 1 and 3 give the same product mixture is attributed to the formation of a common intermediate, i.e., the coordinatively unsaturated 16e species {IrH2(TpMe2)}.  相似文献   

2.
Condensation of X-benzaldehyde with (R)-2-amino-2-phenylethanol gives the enantiopure Schiff bases (R)-2-(X-benzaldimine)-2-phenylethanol (X?=?H, HL1; 2,4-dimethoxy, HL2). The Schiff bases coordinate with dinuclear [Rh(η4-cod)(µ-O2CCH3)]2 to afford the cationic complexes [Rh(η4-cod){(R)-2-(benzaldimine)-2-phenylethanol-κ 2 N,O}](acetate), [Rh(η4-cod)(HL1)](ac) (1) and [Rh(η4-cod){(R)-2-(2,4-dimethoxy-benzaldimine)-2-phenylethanol-κ 2 N,O}](acetate), [Rh(η4-cod)(HL2)](ac) (2), respectively. The Schiff bases and complexes are isolated as solids in good yields and characterized by elemental analysis, IR, UV-Vis, 1H/13C-NMR, mass spectroscopy, and polarimetry.  相似文献   

3.
Condensation of salicyldehyde with (R or S)-2-amino-2-phenylethanol or rac-2-amino-1-phenylethanol gives enantiopure (R or S)-2-(salicylaldimine)-2-phenylethanol (R- or S-H2L1) or (rac)-2-(salicylaldimine)-1-phenylethanol (rac-H2L2). The Schiff bases coordinate to [Rh(η 4-cod)(μ-O2CCH3)]2 to afford mononuclear [Rh(η 4-cod){(R or S)-2-(salicylaldiminato)-2-phenylethanol-κ 2 N,O}], [Rh(η 4-cod)(R- or S-HL1)] (1 or 2), or [Rh(η 4-cod){(rac)-2-(salicylaldiminato)-1-phenylethanol-κ 2 N,O}], [Rh(η 4-cod)(rac-HL2)] (3). The Schiff base and complexes are characterized by IR-, UV/Vis-, 1H/13C-NMR-, mass-spectroscopy, circular dichroism (CD), and polarimetry. The synthetic and spectroscopic results suggest that deprotonated Schiff base coordinates to [Rh(η 4-cod)] as a six-membered N,O-chelate with distorted square planar geometry at rhodium. CD and polarimetry measurements show the enantiopurity of the Schiff bases as well as the complexes in solution. The in situ system composed of [Rh(η 4-cod)Cl]2 and S-H2L1 has been used as a catalyst for the reduction of acetophenone into rac-1-phenylethanol with 85% conversion in diphenylsilane at 0–5°C.  相似文献   

4.
Syntheses of [M(P(O)(OMe)2)(P(OMe)3)4] (M = Co, Rh, Ir) are reported and variable temperature 31P NMR studies on the iridium complex are described. Hydrogen reacts differently with the Ir and Rh complexes giving (IrH2(P(O)(OMe)2)(P(OMe)3)3] and [RhH(P(OMe3)4], respectively. The crystal and molecular structure of the novel compound [IrH(P(OMe)3)4(P(OMe)2OSnMe2Cl2)][SnCl3Me2] is described.  相似文献   

5.
Reaction of [Rh(η4-cod)(S)-amino-acidato] ((S)-amino acidate?=?(S)-O2C-CHR-NH2; cod?=?cycloocta-1,5-diene) with 1,2-bis(diphenylphosphino)ethane (dppe) affords the ionic [Rh(dppe)2]{(S)-O2C-CHR-NH2} (R?=?Me, I; Ph, II) complexes. Reactions with 1,3-bis(diphenylphosphino)propane (dppp) or 2,2,2-tris(diphenylphosphinomethyl)ethane (triphos) give the neutral [Rh(dppp){(S)-O2C-CHR-NH2}] (R?=?Me, III; Ph, IV) or [Rh(η2-triphos){(S)-O2C-CHR-NH2}] (R?=?Me, V; Ph, VI) complexes. The complexes are characterized by elemental analysis, UV–Vis-, IR-, 1H/31P{1H} NMR- and mass-spectroscopy. Two molecules of dppe coordinate to the Rh(I) symmetrically by replacing both cod and (S)-amino acidate to give III. Only one molecule of dppp (or triphos) coordinate to the Rh(I) asymmetrically by replacing only cod to give IIIVI. Two diastereomeric Rh(I)-complexes are present in V and VI. The results further suggest that the ligands are arranged in a distorted square planar geometry around the Rh(I) centre. The use of triphos instead of dppe or dppp yields the same coordination sphere.  相似文献   

6.
Unusual chemical transformations such as three‐component combination and ring‐opening of N‐heterocycles or formation of a carbon–carbon double bond through multiple C–H activation were observed in the reactions of TpMe2‐supported yttrium alkyl complexes with aromatic N‐heterocycles. The scorpionate‐anchored yttrium dialkyl complex [TpMe2Y(CH2Ph)2(THF)] reacted with 1‐methylimidazole in 1:2 molar ratio to give a rare hexanuclear 24‐membered rare‐earth metallomacrocyclic compound [TpMe2Y(μN,C‐Im)(η2N,C‐Im)]6 ( 1 ; Im=1‐methylimidazolyl) through two kinds of C–H activations at the C2‐ and C5‐positions of the imidazole ring. However, [TpMe2Y(CH2Ph)2(THF)] reacted with two equivalents of 1‐methylbenzimidazole to afford a C–C coupling/ring‐opening/C–C coupling product [TpMe2Y{η3‐(N,N,N)‐N(CH3)C6H4NHCH?C(Ph)CN(CH3)C6H4NH}] ( 2 ). Further investigations indicated that [TpMe2Y(CH2Ph)2(THF)] reacted with benzothiazole in 1:1 or 1:2 molar ratio to produce a C–C coupling/ring‐opening product {(TpMe2)Y[μ‐η21‐SC6H4N(CH?CHPh)](THF)}2 ( 3 ). Moreover, the mixed TpMe2/Cp yttrium monoalkyl complex [(TpMe2)CpYCH2Ph(THF)] reacted with two equivalents of 1‐methylimidazole in THF at room temperature to afford a trinuclear yttrium complex [TpMe2CpY(μ‐N,C‐Im)]3 ( 5 ), whereas when the above reaction was carried out at 55 °C for two days, two structurally characterized metal complexes [TpMe2Y(Im‐TpMe2)] ( 7 ; Im‐TpMe2=1‐methyl‐imidazolyl‐TpMe2) and [Cp3Y(HIm)] ( 8 ; HIm=1‐methylimidazole) were obtained in 26 and 17 % isolated yields, respectively, accompanied by some unidentified materials. The formation of 7 reveals an uncommon example of construction of a C?C bond through multiple C–H activations.  相似文献   

7.
A new method for the modification of a silylamino ligand has been developed through mono and dual C(sp3)−H/Si−H cross-dehydrocoupling with silanes. The reaction of [LY{η2-(C,N)-CH2Si(Me2)NSiMe3}] (L=bis(2,6-diisopropylphenyl)-β-diketiminato, L′ ( 1L ′); L=tris(3,5-dimethylpyrazolyl)borate, TpMe2 ( 1TpMe2 )) with 2 equivalents of PhSiH3 in toluene gave the complexes [LY{η2-(C,N)-C(SiH2Ph)2Si(Me2)NSiMe3}] (L=L′ ( 2L’ ); L=TpMe2 ( 2TpMe2 )). Moreover, 1TpMe2 reacted with the secondary silanes Ph2SiH2 and Et2SiH2 to afford the corresponding mono C−H activation products [TpMe2Y{η2-(C,N)-CH(SiHR2)Si(Me2)NSiMe3}] (R=Ph ( 4 b ); R=Et ( 4 c )). The equimolar reaction of 1TpMe2 with PhSiH3 also produced the mono C−H activation product 4 a ([TpMe2Y{η2-(C,N)-CH(SiH2Ph)Si(Me2)NSiMe3}(thf)]). A study of their reactivity showed that 4 a facilely reacted with 2 equivalents of benzothiazole by an unusual 1,1-addition of the C=N bond of the benzothiazolyl unit to the Si−H bond to give the C−H/Si−H cross-dehydrocoupling product [(TpMe2)Y{η3-(N,N,N)-N(SiMe3)SiMe2CH2Si(Ph)(CSC6H4N)(CHSC6H4N)}] ( 5 ). These results indicate that this modification endows the silylamino ligand with novel reactivity.  相似文献   

8.
The Rh1(diolefin)complexes [Rh(nbd)( 2 )][PF6] [Rh(1,5-cod)( 2 )][PF6], and [Rh((Z)-α -acetamidocinnamic acid)( 2 )][PF6] ( 2 = the chiral P,N-ligand (S)-1-[bis(p-methylphenyl)phospino]-2-[p-methoxybenzyl)amino]-3-methylbutane have been prepared and characterized. These complexes exit as a mixture of isomers arising from different five-membered-ring conformations and diastereoisomers due to both the prochiral nitrogen and olefin ligands. The three-dimensional solutions structures of these complexes have been studied with the specific aim of understanding how the chiral pocket is built. Aspects of the exchange dynamics and their possible relevance to homogeneous hydrogenation are discussed The solid-state structure for the nbd complex, [Rh(nbd)( 2 )][PF6], as well as detailed one- and two-dimensional 31P-, 13C-, and 1H-NMR results are presented.  相似文献   

9.
The reactions of the organometallic 1,4-diazabutadienes, RN=C(R′)C(Me)=NR″ [R = R″ = p-C6H4OMe, R′ = trans-PdCl(PPh3)2 (DAB); R = p-C6H4OMe, R″ = Me, R′ = trans-PdCl(PPh3)2 (DABI; R = R″ = p-C6H4OMe, R′ = Pd(dmtc)-(PPh3), dmtc = dimethyldithiocarbamate (DABII); R = R″ = p-C6H4OMe, R′ = PdCl(diphos), diphos = 1,2-bis(diphenylphosphino)ethane (DABIII)] with [RhCl(COD)]2 (COD = 1,5-cyclooctadiene, Pd/Rh ratio = 12) depend on the nature of the ancillary ligands at the Pd atom in group R′. In the reactions with DAB and DABI transfer of one PPh3 ligand from Pd to Rh occurs yielding [RhCl(COD)(PPh3)] and the new binuclear complexes [Rh(COD) {RN=C(R?)-C(Me)=NR″}], in which the diazabutadiene moiety acts as a chelating bidentate ligand. Exchange of ligands between the two different metallic centers also occurs in the reaction with DABII. In this case, the migration of the bidentate dmtc anion yields [Rh(COD)Pdmtc] and [Rh(COD) {RN=C(R?)C(Me)=NR″}]. In contrast, the reaction with DABIII leads to the ionic product [Rh(COD)- (DABIII)][RhCl2(COD)], with no transfer of ligands. The cationic complex [Rh(COD)(DABIII)]+ can be isolated as the perchlorate salt from the same reaction (Pd/Rh ratio = 1/1) in the presence of an excess of NaClO4. In all the binuclear complexes the coordinated 1,5-cyclooctadiene can be readily displaced by carbon monoxide to give the corresponding dicarbonyl derivatives. The reaction of [RhCl(CO)2]2 with DAB and/or DABI yields trinuclear complexes of the type [RhCl(CO)2]2(DAB), in which the diazabutadiene group acts as a bridging bidentate ligand. Some reactions of the organic diazabutadiene RN=C(Me)C(Me)=NR (R = p-C6H4OMe) are also reported for comparison.  相似文献   

10.
The reaction of cationic diolefinic rhodium(I) complexes with 2‐(diphenylphosphino)benzaldehyde (pCHO) was studied. [Rh(cod)2]ClO4 (cod=cycloocta‐1,5‐diene) reacted with pCHO to undergo the oxidative addition of one pCHO with (1,2,3‐η)cyclooct‐2‐en‐1‐yl (η3‐C8H13) formation, and the coordination of a second pCHO molecule as (phosphino‐κP)aldehyde‐κO(σ‐coordination) chelate to give the 18e acyl(allyl)rhodium(III) species [Rh(η3‐C8H13)(pCO)(pCHO)]ClO4 (see 1 ). Complex 1 reacted with [Rh(cod)(PR3)2]ClO4 (R=aryl) derivatives 3 – 6 to give stable pentacoordinated 16e acyl[(1,2,3‐η)‐cyclooct‐2‐en‐1‐yl]rhodium(III) species [Rh(η3‐C8H13)(pCO)(PR3)]ClO4 7 – 10 . The (1,2,3‐η)‐cyclooct‐2‐en‐1‐yl complexes contain cis‐positioned P‐atoms and were fully characterized by NMR, and the molecular structure of 1 was determined by X‐ray crystal diffraction. The rhodium(III) complex 1 catalyzed the hydroformylation of hex‐1‐ene and produced 98% of aldehydes (n/iso=2.6).  相似文献   

11.
Organorhodium complexes, such as RhH(PPh3)4, RhH(CO)(PPh3)3, Rh(η3-C3H4Ph)(CO)(PPh3)2, and RhH(dppe)2 [dppe = 1,2-bis(diphenylphosphinoethane)], catalyze polymerization of phenylallene and of 4-methylphenylallene at 60 °C. High-molecular-weight polymers (Mn>4×105) are isolated from the reaction products by removing the low-molecular-weight (Mn<3×103) acetone-soluble fraction. The NMR (1H and 13C{1H}) spectra of poly(phenylallene) (1) and poly(4-methylphenylallene) (2) show the structure formed through selective 2,3-polymerization of the monomers, while similarly obtained poly(2-naphthylallene) (3) is characterized only by 1H NMR spectroscopy due to its low solubility in common organic solvents. 4-Fluorophenylallene and 4-(trifluoromethyl)-phenylallene do not polymerize under similar conditions in the presence of RhH(PPh3)4 catalyst but are turned into low-molecular-weight oligomers. CoH(N2)(PPh3)3-catalyzed polymerization of phenylallene and 4-methylphenylallene at room temperature gives the corresponding polymers with molecular weights in the range Mn=(9–15)×104, in high yields. © 1997 John Wiley & Sons, Ltd.  相似文献   

12.
Syntheses of [M(P(O)(OMe)2)(P(OMe)3)4] (M = Co, Rh, Ir) and variable temperature 31P NMR studies are described, and mechanistic implications discussed.  相似文献   

13.
The water-soluble Rh(I)-THP complexes: RhCl(1,5-cod)(THP) (), [Rh(1,5-cod)(THP)(2)]Cl (), RhCl(THP)(4) (), and trans-RhCl(CO)(THP)(2) () have been synthesized and characterized, where THP = P(CH(2)OH)(3); - are the first potentially useful entries into Rh(I)-THP chemistry, while and are the first structurally characterized Rh(I)-THP complexes.  相似文献   

14.
The novel cis-(σ-alkyl)(η2-O2) complexes of rhodium [(THF)(EtOH)Naμ-EtOH2μ-(CO2R)CH2CH(CO2R)Rh(η2-O2)(triphos)2Na(EtOH)(THF)][BPh4]2·2EtOH (R = Me,3; Et,4) have been synthesized by reaction of dioxygen with the hydrides (triphos)(RhH(η2-alkene) followed by NaBPh4 addition (alkene = dimethyl fumarate,1; diethyl fumarate,2) (triphos = MeC(CH2PPh2)3). The structure of4 has been determined by X-ray diffraction. Oxygen atom transfer reactions from the η2-O2 complexes to various inorganic and organic substrates have been studied.  相似文献   

15.
Synthesis of Carboxylate Substituted Rhenium Gold Metallatetrahedranes Re2(AuPPh3)2(μ-PCy2)(CO)71-OC(R)O) (R = H, Me, CF3, Ph, 3,4-(OMe)2C6H3) The reaction of the in situ prepared salt Li[Re2(μ-H)(μ-PCy2)(CO)7(ax-C(Ph)O)] ( 2 ) with 1,5 equivalents of monocarboxylic acid RCOOH (R = H ( 4 a ), Me ( 4 b ), CF3 ( 4 c ), Ph ( 4 d ), 3,4-(OMe)2C6H3 ( 4 e ) in tetrahydrofruan (THF) solution at 60 °C gives within 4 h under release of benzaldehyde (PhCHO) the η1-carboxylate substituted dirhenium salt Li[Re2(μ-H)(μ-PCy2)(CO)71-OC(R)O)] (R = H ( 4 a ), Me ( 4 b ), CF3 ( 4 c ), Ph ( 4 d ), 3,4-(OMe)2C6H3 ( 4 e )) in almost quantitative yield. The lower the pKa value of the respective carboxylic acid the faster the reaction proceeds. It was only in the case of CF3COOH possible to prove the formation of the hydroxycarbene complex Re2(μ-H)(μ-PCy2)(CO)7(=C(Ph)OH) ( 5 ) prior to elimination of PhCHO. The new compounds 4 a–4 e were only characterized by 31P NMR and ν(CO) IR spectroscopy as they are only stable in solution. They are converted with two equivalents of BF4AuPPh3 at 0 °C in a so-called cluster expansion reaction into the heterometallic metallatetrahedrane complexes Re2(AuPPh3)2(μ-PCy2)(CO)71-OC(R)O) (R = H ( 7 a ), Me ( 7 b ), CF3 ( 7 c ), Ph ( 7 d ), 3,4-(OMe)2C6H3 ( 7 e )) (yield 47–71% ). The expected precursor complexes of 7 a–7 e Li[Re2(AuPPh3)(μ-PCy2)(CO)71-OC(R)O] ( 8 ) were not detected by NMR and IR spectroscopy in the course of the reaction. Their existence was retrosynthetically proved by the reaction of 7 b with an excess of the chelating base TBD (1,5,7-Triazabicyclo[4.4.0]dec-5-en) forming [(TBD)xAuPPh3][Re2(AuPPh3)(μ-PCy2)(CO)71-OC(Me)O] ( 8 b ) in solution. The η1-bound carboxylate ligand in 7 a–7 e can photochemically be converted into a μ-bound ligand in Re2(AuPPh3)2(μ-PCy2)(μ-OC(R)O)(CO)6 (R = H ( 9 a ), Me ( 9 b ), CF3 ( 9 c ), Ph ( 9 d ), 3.4-(MeO)2C6H3 ( 9 e )) under release of one equivalent CO. All isolated cluster complexes were characterized and identified by the following analytical methods: elementary analysis, NMR (1H, 31P) spectroscopy, ν(CO) IR spectroscopy and in the case of 7 d and 9 b by X-ray structure analysis.  相似文献   

16.
A series of unusual chemical‐bond transformations were observed in the reactions of high active yttrium? dialkyl complexes with unsaturated small molecules. The reaction of scorpionate‐anchored yttrium? dibenzyl complex [TpMe2Y(CH2Ph)2(thf)] ( 1 , TpMe2=tri(3,5‐dimethylpyrazolyl)borate) with phenyl isothiocyanate led to C?S bond cleavage to give a cubane‐type yttrium–sulfur cluster, {TpMe2Y(μ3‐S)}4 ( 2 ), accompanied by the elimination of PhN?C(CH2Ph)2. However, compound 1 reacted with phenyl isocyanate to afford a C(sp3)? H activation product, [TpMe2Y(thf){μ‐η13‐OC(CHPh)NPh}{μ‐η32‐OC(CHPh)NPh}YTpMe2] ( 3 ). Moreover, compound 1 reacted with phenylacetonitrile at room temperature to produce γ‐deprotonation product [(TpMe2)2Y]+[TpMe2Y(N=C?CHPh)3]? ( 6 ), in which the newly formed N?C?CHPh ligands bound to the metal through the terminal nitrogen atoms. When this reaction was carried out in toluene at 120 °C, it gave a tandem γ‐deprotonation/insertion/partial‐TpMe2‐degradation product, [(TpMe2Y)2(μ‐Pz)2{μ‐η13‐NC(CH2Ph)CHPh}] ( 7 , Pz=3,5‐dimethylpyrazolyl).  相似文献   

17.
Summary A synthesis of [RhH{P(OPh)3}4] (1) from [Rh(acac){P(OPh)3}2] or [Rh(acac)(CO)2] has been developed. The reaction of theortho-metallated complex (2) with H2, leading to (1) is described.  相似文献   

18.
A novel, useful in situ synthesis for NHC nickel allyl halide complexes [Ni(NHC)(η3-allyl)(X)] starting from [Ni(CO)4], NHC and allyl halides is presented. The reaction of [Ni(CO)4] with (i) one equivalent of the corresponding NHC and (ii) with an excess of the corresponding allyl chloride at room temperature leads with elimination of carbon monoxide to complexes of the type [Ni(NHC)(η3-allyl)(X)]. This approach was used to synthesize the complexes [Ni(tBu2Im)(η3-H2C -C (Me)-C H2)(Cl)] ( 2 ), [Ni(iPr2ImMe)(η3-H2C -C (Me)-C H2)(Cl)] ( 3 ), [Ni(iPr2Im)(η3-H2C -C (Me)-C H2)(Cl)] ( 4 ), [Ni(iPr2Im)(η3-H2C -C (H)-C (Me)2)(Br)] ( 5 ), [Ni(Me2ImMe)(η3-H2C -C (Me)-C H2)(Cl)] ( 6 ), and [Ni(EtiPrImMe)(η3-H2C -C (Me)-C H2)(Cl)] ( 7 ). The complexes 1 to 7 were characterized using NMR and IR spectroscopy and elemental analysis, and the molecular structures are provided for 2 and 7 . The allyl nickel complexes 1 – 7 are stereochemically non-rigid in solution due to (i) NHC rotation about the nickel-carbon bond, (ii) allyl rotation about the Ni–η3-allyl axis and (iii) π–σ–π allyl isomerization processes. The allyl halide complexes can be methylated as was demonstrated by the methylation of a number of the complexes [Ni(NHC)(η3-allyl)(X)] with methylmagnesium chloride or methyllithium, which led to isolation of the complexes [Ni(Me2Im)(η3-H2C -C (Me)-C H2)(Me)] ( 8 ), [Ni(tBu2Im)(η3-H2C -C (Me)-C H2)(Me)] ( 9 ), [Ni(iPr2ImMe)(η3-H2C -C (Me)-C H2)(Me)] ( 10 ), [Ni(iPr2Im)(η3-H2C -C (Me)-C H2)(Me)] ( 11 ), [Ni(iPr2Im)(η3-H2C -C (H)-C (Me)2)(Me)] ( 12 ), and [Ni(EtiPrImMe)(η3-H2C -C (Me)-C H2)(Me)] ( 13 ). These complexes were fully characterized including X-ray molecular structures for 10 and 11 .  相似文献   

19.
《Polyhedron》2004,23(2-3):379-383
In an effort to obtain suitable starting materials, the synthesis and crystal structure of hydrotris(3,5-dimethylpyrazolyl)borato (TpMe2) tantalum(III) complexes are reported. Reaction of KTpMe2 with [TaCl2(tht)]2(μ-Cl)2(μ-tht) (1) (tht=tetrahydrothiophene) gives the X-ray characterized unsymmetrical (TpMe2Ta)(TaCl3)(μ-Cl2)(μ-tht) (2) in fair yield. This doubly bonded TaTa complex [2.6791(5) Å, diamagnetic, 18e per Ta] is a rare example of an unsymmetrical Ta2X6L3 complex. The X-ray structure of TpMe2TaCl2(PhCCMe) (3) is also reported. It has the (4e)-alkyne in the symmetry plane of the molecule. The trans effects of tht and phenylpropyne are discussed. 1 is inert towards the reaction with a second equivalent of KTpMe2.  相似文献   

20.
The ease of formation of the phosphonate complexes [M(P(O)(OMe)2(P(OMe)3)4] (M = Rh, Ir), from the pentakis-trimethylphosphite complexes [M(P(OMe)3)5]Cl is reported. Differences in the interaction of H2 with the complexes [M′(P(O)(OMe)2(P(OMe)3)4), (M′ = Co, Rh, Ir) are presented and discussed.  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