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.
The synthesis and characterization of a biomimetic system that can reversibly bind acetylene (ethyne) is reported. The system has been designed to mimic catalytic intermediates of the tungstoenzyme acetylene hydratase. The thiophenyloxazoline ligand S‐Phoz (2‐(4′,4′‐dimethyloxazolin‐2′‐yl)thiophenolate) is used to generate a bioinspired donor environment around the W center, facilitating the stabilization of W–acetylene adducts. The featured complexes [W(C2H2)(CO)(S‐Phoz)2] ( 2 ) and [WO(C2H2)(S‐Phoz)2] ( 3 ) are extremely rare from a synthetic and structural point of view as very little is known about W–C2H2 adducts. Upon exposure to visible light, 3 can release C2H2 from its coordination sphere to yield the 14‐electron species [WO(S‐Phoz)2] ( 4 ). Under light‐exclusion 4 re‐activates C2H2 making this the first fully characterized system for the reversible activation of acetylene.  相似文献   

2.
A series of heterodinuclear complexes with acetylene dithiolate (acdt2?) as the bridging moiety were synthesised by a facile one‐pot procedure that avoided use of the highly elusive acetylene dithiol. Generation of the W–Ru complex [Tp′W(CN)(CO)(C2S2)Ru(η5‐C5H5)(PPh3)] (Tp’=hydrotris(3,5‐dimethylpyrazolyl)borate) and the W–Pd complexes [Tp′W(CN)(CO)(C2S2)Pd(dppe)] and [Tp′W(CO)2(C2S2)Pd(dppe)][PF6] (dppe=1,2‐bis(diphenylphoshino)ethane), which exhibit a [W(η2‐κ2‐C2S2)M] core (M=Ru, Pd), was accomplished by using a transition‐metal‐assisted solvolytical removal of the Me3Si‐ethyl thiol protecting groups. All intermediate species of the reaction have been fully characterised. The highly coloured W–Ru complex [Tp′W(CN)(CO)(C2S2)Ru(η5‐C5H5)(PPh3)] shows reversible redox chemistry, as does the prototype complex [Tp′W(CO)2(C2S2)Ru(η5‐C5H5)(PPh3)][PF6]. Single crystal X‐ray diffraction and IR, EPR and UV/Vis spectroscopic studies in conjunction with DFT calculations prove the high electronic delocalisation of states over the acdt2? linker. Comparative studies revealed a higher donor strength and more pronounced dithiolate character of acdt2? in [Tp′W(CN)(CO)(C2S2)Ru(η5‐C5H5)(PPh3)] relative to [Tp′W(CO)2(C2S2)Ru(η5‐C5H5)(PPh3)]+. In addition, the influence of the overall complex charge on the metric parameters was investigated by single‐crystal X‐ray diffraction studies with the W–Pd complexes [Tp′WL2(C2S2)Pd(dppe)] (L=(CN?)(CO) or (CO)2). The central [W(C2S2)Pd] units exhibit high structural similarity, which indicates the extensive delocalisation of charge over both metals.  相似文献   

3.
A series of WIV alkyne complexes with the sulfur-rich ligand hydridotris(2-mercapto-1-methylimidazolyl) borate) (TmMe) are presented as bio-inspired models to elucidate the mechanism of the tungstoenzyme acetylene hydratase (AH). The mono- and/or bis-alkyne precursors were reacted with NaTmMe and the resulting complexes [W(CO)(C2R2)(TmMe)Br] (R=H 1 , Me 2 ) oxidized to the target [WE(C2R2)(TmMe)Br] (E=O, R=H 4 , Me 5 ; E=S, R=H 6 , Me 7 ) using pyridine-N-oxide and methylthiirane. Halide abstraction with TlOTf in MeCN gave the cationic complexes [WE(C2R2)(MeCN)(TmMe)](OTf) (E=CO, R=H 10 , Me 11 ; E=O, R=H 12 , Me 13 ; E=S, R=H 14 , Me 15 ). Without MeCN, dinuclear complexes [W2O(μ-O)(C2Me2)2(TmMe)2](OTf)2 ( 8 ) and [W2(μ-S)2(C2Me2)(TmMe)2](OTf)2 ( 9 ) could be isolated showing distinct differences between the oxido and sulfido system with the latter exhibiting only one molecule of C2Me2. This provides evidence that a fine balance of the softness at W is important for acetylene coordination. Upon dissolving complex 8 in acetonitrile complex 13 is reconstituted in contrast to 9 . All complexes exhibit the desired stability toward water and the observed effective coordination of the scorpionate ligand avoids decomposition to disulfide, an often-occurring reaction in sulfur ligand chemistry. Hence, the data presented here point toward a mechanism with a direct coordination of acetylene in the active site and provide the basis for further model chemistry for acetylene hydratase.  相似文献   

4.
The feasibility of oxidative addition of the P−H bond of PHPh2 to a series of rhodium complexes to give mononuclear hydrido-phosphanido complexes has been analyzed. Three main scenarios have been found depending on the nature of the L ligand added to [Rh(Tp)(C2H4)(PHPh2)] (Tp= hydridotris(pyrazolyl)borate): i) clean and quantitative reactions to terminal hydrido-phosphanido complexes [RhTp(H)(PPh2)(L)] (L=PMe3, PMe2Ph and PHPh2), ii) equilibria between RhI and RhIII species: [RhTp(H)(PPh2)(L)]⇄[RhTp(PHPh2)(L)] (L=PMePh2, PPh3) and iii) a simple ethylene replacement to give the rhodium(I) complexes [Rh(κ2-Tp)(L)(PHPh2)] (L=NHCs-type ligands). The position of the P−H oxidative addition–reductive elimination equilibrium is mainly determined by sterics influencing the entropy contribution of the reaction. When ethylene was used as a ligand, the unique rhodaphosphacyclobutane complex [Rh(Tp)(η1-Et)(κC,P-CH2CH2PPh2)] was obtained. DFT calculations revealed that the reaction proceeds through the rate limiting oxidative addition of the P−H bond, followed by a low-barrier sequence of reaction steps involving ethylene insertion into the Rh−H and Rh−P bonds. In addition, oxidative addition of the P−H bond in OPHPh2 to [Rh(Tp)(C2H4)(PHPh2)] gave the related hydride complex [RhTp(H)(PHPh2)(POPh2)], but ethyl complexes resulted from hydride insertion into the Rh−ethylene bond in the reaction with [Rh(Tp)(C2H4)2].  相似文献   

5.
The oxidative cleavage of [Fe2(η-C5H5)2(CO)4-n(CNMe)n] (n=0−2) by 2AgX gives mononuclear products. It is shown to be a two-electron process in most solvents but a one-electron process in acetonitrile. The two-electron oxidations proceed by way of adducts such as [Fe2(η-C5H5)2(CO)(CNMe)(μ-CO){;μ-CN(Me)AgPPh3};]BF4 which are isolable when n = 2, detectable when n = 1 and postulatetd when n = 0. The one-electron process gives no adducts, and 1AgX cleaves all of the substrate to [Fe(η-C5H5)(CO)(L)(NCMe)]+ and [Fe(η-C5H5)(CO)(L)]. (L  CO or CNME). The latter may combine or react with added CHBr3 to give [Fe(η-C5H5)(CO)(L)Br]. The structure of [Fe(η-C5H5)(CO)2-(CNMe)]BF4 has been determined by X-ray diffraction.  相似文献   

6.
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   

7.
In the course of the study of the reactivity of the dinuclear complexes [RhCl(CO)(C2H2)]2 and [RhSR(CO)2]2 towards nucleophiles, two series of dinuclear pentacoordinated rhodium(I) complexes, [RhCl(CO)(C2H4)(amine)]2 and [RhSR(CO)2PR3]2, have been isolated.  相似文献   

8.
The effect of solvation on the conformation of acetylene has been studied by adding one water molecule at a time. Quantum chemical calculations of the H+(C2H2)(H2O)n (n=1-5) clusters indicate that the H2O molecules prefer to form the OH…π interaction rather than the CH…O interaction. This solvation motif is different from that of neutral (C2H2)(H2O)n (n=1-4) clusters, in which the H2O molecules prefer to form the CH…O and OH…C H-bonds. For the H+(C2H2)(H2O)n cationic clusters, the first solvation shell consists of one ring structure with two OH…π H-bonds and three water molecules, which is completed at n=4. Simulated infrared spectra reveal that vibrational frequencies of OH…π H-bonded O-H stretching afford a sensitive probe for exploring the solvation of acetylene by protonated water molecules. Infrared spectra of the H+(C2H2)(H2O)n(n=1-5) clusters could be readily measured by the infrared photodissociation technique and thus provide useful information for the understanding of solvation processes.  相似文献   

9.
2 a and 2 b , [Ir(CI)(COD)(NHC)] (COD=1,5-cyclooctadiene), have been prepared via transmetallation from NHC−Ag complexes. [Rh(CI)(COD)(NHC)] ( 4 ) was prepared analogously. [Ir({κ-C,N-(NHC-acetamide−1H)}(COD)] ( 3 c ) has been synthesized via transmetallation from the deprotonated NHC−Ag complex. [IrCp*({κ-C,N-(NHC-acetamide−1H)}] ( 5 ) (Cp*=pentamethylcyclopentadienyl), has been obtained analogously. [Ir(CI)(CO)2(NHC)] ( 6 ) and [Ir({κ-C,N-(NHC-acetamide−1H)}(CO)2] ( 7 ) have been prepared by carbonylation of 2 b and 3 c , respectively. The catalytic activity of these complexes has been evaluated in the dehydrogenation of formic acid, under solventless conditions, in the presence of water as a cosolvent, and in a 5 : 2 HCOOH/Et3N mixture, with the best TOF values being obtained in the case of the latter. Stoichiometric experiments suggest COD hydrogenation as the preactivation step.  相似文献   

10.
The two‐step one‐pot oxidative decarbonylation of [Fe2(S2C2H4)(CO)4(PMe3)2] ( 1 ) with [FeCp2]PF6, followed by addition of phosphane ligands, led to a series of diferrous dithiolato carbonyls 2 – 6 , containing three or four phosphane ligands. In situ measurements indicate efficient formation of 1 2+ as the initial intermediate of the oxidation of 1 , even when a deficiency of the oxidant was employed. Subsequent addition of PR3 gave rise to [Fe2(S2C2H4)(μ‐CO)(CO)3(PMe3)3]2+ ( 2 ) and [Fe2(S2C2H4)(μ‐CO)(CO)2(PMe3)2(PR3)2]2+ (R=Me 3 , OMe 4 ) as principal products. One terminal CO ligand in these complexes was readily substituted by MeCN, and [Fe2(S2C2H4)(μ‐CO)(CO)2(PMe3)3(MeCN)]2+ ( 5 ) and [Fe2(S2C2H4)(μ‐CO)(CO)(PMe3)4(MeCN)]2+ ( 6 ) were fully characterized. Relevant to the Hred state of the active site of Fe‐only hydrogenases, the unsymmetrical derivatives 5 and 6 feature a semibridging CO ligand trans to a labile coordination site.  相似文献   

11.
Treatment of the (isocyanide)gold(I) species LAuCl (L=tBuNC, 2,6‐Me2C6H3NC) with 4‐mercaptobenzoic acid in the presence of NaOMe yields the complexes [Au(4‐SC6H4CO2H)L] in good yield. Reaction of LAuCl with 2‐HSQn (Qn=quinoline) and 2‐HSPy (Py=pyridine) under the same conditions provides the thiolato compounds [Au(2‐SQn)L] and [Au(2‐SPy)L], respectively. A structural investigation of the pyridylthiolato compound revealed chains of molecules with alternating medium and long Au−Au interactions. Treatment of this compound with HBF4 results in the cationic species [Au(2‐HSPy)(2,6‐Me2C6H3NC)]+ as the BF4 salt. The same product is obtained on reaction of [AuCl(2,6‐Me2C6H3NC)] with AgOTf followed by HSPy. Treatment of the gold(I) halide compounds LAuCl (L=tBuNC, 2,6‐Me2C6H3NC) with potassium 1,3,4‐thiadiazole‐2,5‐dithiolate (KSSSK) leads to the isolation of dinuclear thiolatogold complexes [(AuL)2(SSS)]. These products go on to form insoluble polymers through loss of isocyanide on standing in solution. A single crystal of [{Au(tBuNC)}2(SSS)] was obtained and the subsequent structural analysis revealed one of the most complicated networks known based solely on aurophilic interactions. A good comparison to the ‘soft' S‐donation of the thiolato ligands was provided by the synthesis of a number of nitratogold(I)complexes with the anion bound through the ‘hard' O‐donor. Reaction of iPrNC and CyNC with Au(tht)Cl provided the complexes [AuCl(iPrNC)] and [AuCl(CyNC)], respectively. These compounds were found to yield the respective nitrato species [Au(NO3)iPrNC)] and [(Au(NO3)(CyNC)] on treatment with AgNO3. The nitrato complexes yielded single crystals enabling a structural investigation to be carried out. While [Au(NO3)(CyNC)] has a more conventional structure with dimers aligned into strings with alternating short and long aurophilic bonding, [Au(NO3)(iPrNC)] has a unique structure based on strings of alternating, corner‐sharing Au6 and Au8 units with short Au−Au contacts in edge‐sharing Au3 triangles.  相似文献   

12.
In the ion/molecule reactions of the cyclometalated platinum complexes [Pt(L? H)]+ (L=2,2′‐bipyridine (bipy), 2‐phenylpyridine (phpy), and 7,8‐benzoquinoline (bq)) with linear and branched alkanes CnH2n+2 (n=2–4), the main reaction channels correspond to the eliminations of dihydrogen and the respective alkenes in varying ratios. For all three couples [Pt(L? H)]+/C2H6, loss of C2H4 dominates clearly over H2 elimination; however, the mechanisms significantly differs for the reactions of the “rollover”‐cyclometalated bipy complex and the classically cyclometalated phpy and bq complexes. While double hydrogen‐atom transfer from C2H6 to [Pt(bipy? H)]+, followed by ring rotation, gives rise to the formation of [Pt(H)(bipy)]+, for the phpy and bq complexes [Pt(L? H)]+, the cyclometalated motif is conserved; rather, according to DFT calculations, formation of [Pt(L? H)(H2)]+ as the ionic product accounts for C2H4 liberation. In the latter process, [Pt(L? H)(H2)(C2H4)]+ (that carries H2 trans to the nitrogen atom of the heterocyclic ligand) serves, according to DFT calculation, as a precursor from which, due to the electronic peculiarities of the cyclometalated ligand, C2H4 rather than H2 is ejected. For both product‐ion types, [Pt(H)(bipy)]+ and [Pt(L? H)(H2)]+ (L=phpy, bq), H2 loss to close a catalytic dehydrogenation cycle is feasible. In the reactions of [Pt(bipy? H)]+ with the higher alkanes CnH2n+2 (n=3, 4), H2 elimination dominates over alkene formation; most probably, this observation is a consequence of the generation of allyl complexes, such as [Pt(C3H5)(bipy)]+. In the reactions of [Pt(L? H)]+ (L=phpy, bq) with propane and n‐butane, the losses of the alkenes and dihydrogen are of comparable intensities. While in the reactions of “rollover”‐cyclometalated [Pt(bipy? H)]+ with CnH2n+2 (n=2–4) less than 15 % of the generated product ions are formed by C? C bond‐cleavage processes, this value is about 60 % for the reaction with neo‐pentane. The result that C? C bond cleavage gains in importance for this substrate is a consequence of the fact that 1,2‐elimination of two hydrogen atoms is no option; this observation may suggest that in the reactions with the smaller alkanes, 1,1‐ and 1,3‐elimination pathways are only of minor importance.  相似文献   

13.
Dinuclear rhodium complexes of the type [Rh2(C2O4)(diolefin)2] (diolefin)2  1,5-cyclooctadiene, 2,5-norbornadiene and tetrafluorobenzobarrelene) with bridging oxalate ligands have been obtained by reaction of [Rh(acac)(diolefin)] with oxalic acid (2: 1 mol ratio). The use of a 1 : 1 molar ratio affords [Rh(HC2O4)(COD)], that reacts with [Ir(acac)(COD)] yielding the heterodinuclear [(COD)Rh(C2O4)Ir(COD)] complex. Treatment of [Rh2(C2O4)(diolefin)2] complexes with phenanthroline type ligands leads to ionic complexes of formula [Rh(diolefin) (phen)][Rh(C2O4)(diolefin)]. Bubbling of carbon monoxide through solutions of the diolefin complexes leads to the formation of carbonylrhodium species of formula [Rh2(C2O4)(CO)2L2] (L = CO, PPh3t-BuNC) or [Rh(CO)2(phen)] - [Rh(C2O4)(CO)2]. Other related malonate complexes are also described.  相似文献   

14.
The “CPNR” ligand may be viewed as being isolobal with fulminate, CNO; however, attempts to prepare a complex of such a ligand resulted instead in a range of novel imino and aminophosphinocarbyne complexes. Sequential treatment of [Mo(≡CBr)(CO)2(Tp*)] (Tp*=hydrotris(dimethylpyrazolyl)borate) with nBuLi and ClP=NMes* (Mes*=C6H2tBu3-2,4,6) afforded mixtures of the complexes [Mo(≡CPnBuNHMes*)(CO)2(Tp*)] and traces of the bimetallic products [Mo22-C2P2O(NHMes)2}(CO)4(Tp*)2] and [Mo22-C2PNHMes)(CO)4(Tp*)2]. The reaction of [W(≡CBr)(CO)2(Tp*)] with nBuLi and ClP=NMes* afforded predominantly the mononuclear carbyne [W{≡CP(=NMes*)nBu2})(CO)2(Tp*)] and traces of the binuclear complex [W2(μ-C2PNHMes)(CO)4(Tp*)2] which is also obtained when tBuLi is used. Although not isolable, the intended complexes [M(≡CPNMes*)(CO)2(Tp*)] could be generated in situ and spectroscopically characterized via the reactions of the stannyl carbynes [M(≡CSnnBu3)(CO)2(Tp*)] and ClP=NMes*. The preceding observations are mechanistically interpreted with reference to a computational interrogation of the model complex [Mo(≡CP=NCH3)(CO)2(Tp*)], the LUMO of which has considerable phosphorus character.  相似文献   

15.
The compounds tricarbonyl(η5‐1‐iodocyclopentadienyl)manganese(I), [Mn(C5H4I)(CO)3], (I), and tricarbonyl(η5‐1‐iodocyclopentadienyl)rhenium(I), [Re(C5H4I)(CO)3], (III), are isostructural and isomorphous. The compounds [μ‐1,2(η5)‐acetylenedicyclopentadienyl]bis[tricarbonylmanganese(I)] or bis(cymantrenyl)acetylene, [Mn2(C12H8)(CO)6], (II), and [μ‐1,2(η5)‐acetylenedicyclopentadienyl]bis[tricarbonylrhenium(I)], [Re2(C12H8)(CO)6], (IV), are isostructural and isomorphous, and their molecules display inversion symmetry about the mid‐point of the ligand C[triple‐bond]C bond, with the (CO)3M(C5H4) (M = Mn and Re) moieties adopting a transoid conformation. The molecules in all four compounds form zigzag chains due to the formation of strong attractive I...O [in (I) and (III)] or π(CO)–π(CO) [in (I) and (IV)] interactions along the crystallographic b axis. The zigzag chains are bound to each other by weak intermolecular C—H...O hydrogen bonds for (I) and (III), while for (II) and (IV) the chains are bound to each other by a combination of weak C—H...O hydrogen bonds and π(Csp2)–π(Csp2) stacking interactions between pairs of molecules. The π(CO)–π(CO) contacts in (II) and (IV) between carbonyl groups of neighboring molecules, forming pairwise interactions in a sheared antiparallel dimer motif, are encountered in only 35% of all carbonyl interactions for transition metal–carbonyl compounds.  相似文献   

16.
The reactions of N-methyl-ortho-phenylenediamine (1) with K2MCl4 (M = Pd or Pt) were investigated. At the first stage, the tetraamine dicationic complexes [(C6H4(NH2)(NMeH))2M]Cl2 were formed. Oxidation of these compounds with atmospheric oxygen in an alkaline medium afforded the neutral semiquinonediimine complexes [C6H4(NH)(NMe)]2M. The structures of the {[(C6H4)(NH2)(NHMe)]2Pd}Cl2 and [C6H4(NH)(NPh)]2Pt complexes were established by X-ray diffraction analysis.  相似文献   

17.
Dimeric chlorobridge complex [Rh(CO)2Cl]2 reacts with two equivalents of a series of unsymmetrical phosphine–phosphine monoselenide ligands, Ph2P(CH2)nP(Se)Ph2 {n = 1( a ), 2( b ), 3( c ), 4( d )}to form chelate complex [Rh(CO)Cl(P∩Se)] ( 1a ) {P∩Se = η2‐(P,Se) coordinated} and non‐chelate complexes [Rh(CO)2Cl(P~Se)] ( 1b–d ) {P~Se = η1‐(P) coordinated}. The complexes 1 undergo oxidative addition reactions with different electrophiles such as CH3I, C2H5I, C6H5CH2Cl and I2 to produce Rh(III) complexes of the type [Rh(COR)ClX(P∩Se)] {where R = ? C2H5 ( 2a ), X = I; R = ? CH2C6H5 ( 3a ), X = Cl}, [Rh(CO)ClI2(P∩Se)] ( 4a ), [Rh(CO)(COCH3)ClI(P~Se)] ( 5b–d ), [Rh(CO)(COH5)ClI‐(P~Se)] ( 6b–d ), [Rh(CO)(COCH2C6H5)Cl2(P~Se)] ( 7b–d ) and [Rh(CO)ClI2(P~Se)] ( 8b–d ). The kinetic study of the oxidative addition (OA) reactions of the complexes 1 with CH3I and C2H5I reveals a single stage kinetics. The rate of OA of the complexes varies with the length of the ligand backbone and follows the order 1a > 1b > 1c > 1d . The CH3I reacts with the different complexes at a rate 10–100 times faster than the C2H5I. The catalytic activity of complexes 1b–d for carbonylation of methanol is evaluated and a higher turnover number (TON) is obtained compared with that of the well‐known commercial species [Rh(CO)2I2]?. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

18.
Carbonyl(cycloheptatrienyl)iodo(phosphorus donor)tungstens ([WI(C7H7)(CO)L]; L = P(OMe)3, 1a ; L = P[O(i-Pr)]3, 1b ; L = PPh3, 1c ) were prepared from dicarbonyl(cycloheptatrienyl)iodotungsten ([WI(C7H7)(CO)2)] via a carbonyl-substitution process. Similarly, bromocarbonyl(phosphorus donor)(1,2,4,6-tetramethylcycloheptatrienyl)tungstens ([WBr(Me4C7H3)(CO)L]; L = P(OMe)3, 6a ; L = P[O(i-Pr)]3, 6b ; L = PPh3, 6c ) were obtained from the reaction of bromodicarbonyl(1,2,4,6)-tetramethylcycloheptatrienyl)tungsten ([WBr(Me4C7H3)(CO)2]; 4 ) with L. The reduction of 1a - c , 4 , and 6a , b with sodiumdihydridobis(2-methoxyethoxy)aluminium in toluene led to stable hydrido complexes [WH(R4C7H3)(CO)L] (R = H, L = P(OMe)3, 2a ; R = H, L = P[O(i-Pr)]3, 2b ; R = H, L = PPh3, 2c ; R = Me, L = P(OMe)3, 7a ; R = Me, L = P[O(i-Pr)]3, 7b ; R = Me, L = CO, 7d ). Complexes 2a and 7b were characterized by X-ray structure analyses.  相似文献   

19.
A series of related acetylacetonate–carbonyl–rhodium compounds substituted by functionalized phosphines has been prepared in good to excellent yields by the reaction of [Rh(acac)(CO)2] (acac is acetylacetonate) with the corresponding allyl‐, cyanomethyl‐ or cyanoethyl‐substituted phosphines. All compounds were fully characterized by 31P, 1H, 13C NMR and IR spectroscopy. The X‐ray structures of (acetylacetonato‐κ2O,O′)(tert‐butylphosphanedicarbonitrile‐κP)carbonylrhodium(I), [Rh(C5H7O2)(CO)(C8H13N2)] or [Rh(acac)(CO)(tBuP(CH2CN)2}] ( 2b ), (acetylacetonato‐κ2O,O′)carbonyl[3‐(diphenylphosphanyl)propanenitrile‐κP]rhodium(I), [Rh(C5H7O2)(C15H14N)(CO)] or [Rh(acac)(CO){Ph2P(CH2CH2CN)}] ( 2h ), and (acetylacetonato‐κ2O,O′)carbonyl[3‐(di‐tert‐butylphosphanyl)propanenitrile‐κP]rhodium(I), [Rh(C5H7O2)(C11H22N)(CO)] or [Rh(acac)(CO){tBu2P(CH2CH2CN)}] ( 2i ), showed a square‐planar geometry around the Rh atom with a significant trans influence over the acetylacetonate moiety, evidenced by long Rh—O bond lengths as expected for poor π‐acceptor phosphines. The Rh—P distances displayed an inverse linear dependence with the coupling constants JP‐Rh and the IR ν(C[triple‐bond]O) bands, which accounts for the Rh—P electronic bonding feature (poor π‐acceptors) of these complexes. A combined study from density functional theory (DFT) calculations and an evaluation of the intramolecular H…Rh contacts from X‐ray diffraction data allowed a comparison of the conformational preferences of these complexes in the solid state versus the isolated compounds in the gas phase. For 2b , 2h and 2i , an energy‐framework study evidenced that the crystal structures are mainly governed by dispersive energy. In fact, strong pairwise molecular dispersive interactions are responsible for the columnar arrangement observed in these complexes. A Hirshfeld surface analysis employing three‐dimensional molecular surface contours and two‐dimensional fingerprint plots indicated that the structures are stabilized by H…H, C…H, H…O, H…N and H…Rh intermolecular interactions.  相似文献   

20.
《Comptes Rendus Chimie》2008,11(8):922-925
The hydrogenase model [Fe2(S2C3H6)(CN)2(CO)4]2− was employed as a molecular tecton for the construction of supramolecular aggregates. IR spectroscopy indicated that cyanide bridged aggregates are formed when [Fe2(S2C3H6)(CN)2(CO)4]2− was treated with Lewis acids such as Zn(tetraphenylporphyrinate), [Cu(NCMe)(2,2′-bipyridine)]PF6 and [Cu(NCMe)4]PF6. Condensation of [Fe2(S2C3H6)(CN)2(CO)4]2− with the tritopic Lewis acid [Cp1Rh]2+ afforded the novel expanded tetrahedron cage, {[Fe2(S2C3H6)(CN)2(CO)4]6[Cp1Rh]4}4−. The tetrahedron cage was characterized crystallographically as the PPN salt.  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