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.
A current goal in heterogeneous catalysis is to transfer the synthetic, as well as developing mechanistic, insights from the modern revolution in nanoparticle science to the synthesis of supported-nanoparticle heterogeneous catalysts. In a recent study (Mondloch, J. E.; Wang, Q.; Frenkel, A. I.; Finke, R. G. J. Am. Chem. Soc. 2010, 132, 9701-9714), we initialized tests of the global hypothesis that quantitative kinetic and mechanistic studies, of supported-nanoparticle heterogeneous catalyst formation in contact with solution, can provide synthetic and mechanistic insights that can eventually drive improved syntheses of composition-, size-, and possibly shape-controlled catalysts. That study relied on the development of a well-characterized Ir(1,5-COD)Cl/γ-Al(2)O(3) precatalyst, which, when in contact with solution and H(2), turns into a nonaggregated Ir(0)(~900)/γ-Al(2)O(3) supported-nanoparticle heterogeneous catalyst. The kinetics of the Ir(1,5-COD)Cl/γ-Al(2)O(3) to Ir(0)(~900)/γ-Al(2)O(3) conversion were followed and fit by a two-step mechanism consisting of nucleation (A → B, rate constant k(1)) followed by autocatalytic surface growth (A + B → 2B, rate constant k(2)). However, a crucial, but previously unanswered question is whether the nucleation and growth steps occur primarily in solution, on the support, or possibly in both phases for one or more of the catalyst-formation steps. The present work investigates this central question for the prototype Ir(1,5-COD)Cl/γ-Al(2)O(3) to Ir(0)(~900)/γ-Al(2)O(3) system. Solvent variation-, γ-Al(2)O(3)-, and acetone-dependent kinetic data, along with UV-vis spectroscopic and gas-liquid-chromatography (GLC) data, are consistent with and strongly supportive of a supported-nanoparticle formation mechanism consisting of Ir(1,5-COD)Cl(solvent) dissociation from the γ-Al(2)O(3) support (i.e., from Ir(1,5-COD)Cl/γ-Al(2)O(3)), solution-based nucleation from that dissociated Ir(1,5-COD)Cl(solvent) species, fast Ir(0)(n) nanoparticle capture by γ-Al(2)O(3), and then subsequent solid-oxide-based nanoparticle growth from Ir(0)(n)/γ-Al(2)O(3) and with Ir(1,5-COD)Cl(solvent), the first kinetically documented mechanism of this type. Those data disprove a solid-oxide-based nucleation and growth pathway involving only Ir(1,5-COD)Cl/γ-Al(2)O(3) and also disprove a solution-based nanoparticle growth pathway involving Ir(1,5-COD)Cl(solvent) and Ir(0)(n) in solution. The present mechanistic studies allow comparisons of the Ir(1,5-COD)Cl/γ-Al(2)O(3) to Ir(0)(~900)/γ-Al(2)O(3) supported-nanoparticle formation system to the kinetically and mechanistically well-studied, Ir(1,5-COD)·P(2)W(15)Nb(3)O(62)(8-) to Ir(0)(~300)·(P(2)W(15)Nb(3)O(62)(8-))(n)(-8n) solution-based, polyoxoanion-stabilized nanoparticle formation and stabilization system. That comparison reveals closely analogous, solution Ir(1,5-COD)(+) or Ir(1,5-COD)Cl-mediated, mechanisms of nanoparticle formation. Overall, the hypothesis supported by this work is that these and analogous studies hold promise of providing a way to transfer the synthetic and mechanistic insights, from the modern revolution in nanoparticle synthesis and characterization in solution, to the rational, mechanism-directed syntheses of solid oxide-supported nanoparticle heterogeneous catalysts, also in contact with solution.  相似文献   

2.
Ott LS  Finke RG 《Inorganic chemistry》2006,45(20):8382-8393
The nanocluster literature contains a wide variety of nanocluster stabilizing agents. In addition to the plethora of putative stabilizing additives, 12 claims appear of "solvent-only" stabilization of transition-metal nanoclusters-a hypothesis that is tested for the first time as part of the present studies. When the two main modes of nanocluster stabilization, electrostatic and steric are considered, "solvent-only" stabilization can only be steric (i.e., is not electrostatic). Solvent-only stabilization would, therefore, require that a strongly coordinated, perhaps even kinetically nonlabile, solvent be present on the nanocluster surface. Hence, an investigation has been conducted into potential sources for the stabilization of prototype Ir(0)n transition-metal nanoclusters prepared from [(1,5-COD)Ir(CH3CN)2][BF4] in five different solvents, with a special focus on the formulation and testing of alternative hypotheses regarding the true source of the nanocluster stabilization in putative solvent-only stabilization conditions. Seven total hypotheses are tested with five being initially ruled out; they are, namely, stabilization by (i) trace chloride (ii), surface hydrides, (iii) scavenged charge, (iv) solvent oxidative addition reactions with the nanocluster surface, or (v) polymerized solvent. This led in turn to two additional main alternative hypotheses: (vi) nanocluster surface ligation by high-donor number solvents (i.e., in the absence of anions) and (vii) nanocluster stabilization by surface-coordination of the traditionally weakly coordinating anion BF(4-). Our results reveal a significant contribution to nanocluster stability from the traditionally weakly coordinating BF(4-) in high dielectric constant solvents, such as propylene carbonate. Literature claims of solvent-only nanocluster stabilization are not supported by our findings. Overall, DLVO (Derjaugin-Landau-Verwey-Overbeek) theory of colloidal stability is supported and found to apply to even traditionally weakly coordinating anions.  相似文献   

3.
Reported herein is the synthesis of the previously unknown [Ir(1,5-COD)(μ-H)](4) (where 1,5-COD = 1,5-cyclooctadiene), from commercially available [Ir(1,5-COD)Cl](2) and LiBEt(3)H in the presence of excess 1,5-COD in 78% initial, and 55% recrystallized, yield plus its unequivocal characterization via single-crystal X-ray diffraction (XRD), X-ray absorption fine structure (XAFS) spectroscopy, electrospray/atmospheric pressure chemical ionization mass spectrometry (ESI-MS), and UV-vis, IR, and nuclear magnetic resonance (NMR) spectroscopies. The resultant product parallels--but the successful synthesis is different from, vide infra--that of the known and valuable Rh congener precatalyst and synthon, [Rh(1,5-COD)(μ-H)](4). Extensive characterization reveals that a black crystal of [Ir(1,5-COD)(μ-H)](4) is composed of a distorted tetrahedral, D(2d) symmetry Ir(4) core with two long [2.90728(17) and 2.91138(17) ?] and four short Ir-Ir [2.78680 (12)-2.78798(12) ?] bond distances. One 1,5-COD and two edge-bridging hydrides are bound to each Ir atom; the Ir-H-Ir span the shorter Ir-Ir bond distances. XAFS provides excellent agreement with the XRD-obtained Ir(4)-core structure, results which provide both considerable confidence in the XAFS methodology and set the stage for future XAFS in applications employing this Ir(4)H(4) and related tetranuclear clusters. The [Ir(1,5-COD)(μ-H)](4) complex is of interest for at least five reasons, as detailed in the Conclusions section.  相似文献   

4.
A review of the literature of kinetic and mechanistic studies of transition-metal nanocluster nucleation and growth is presented; the focus is on nucleation processes. A brief survey of nucleation theory is given first, with an emphasis on classical nucleation theory, as this is the logical starting point of transition-metal nanocluster nucleation and growth studies. The main experimental methods for following nanocluster formation are examined next--dynamic light scattering, UV-visible spectroscopy, electron microscopy, and X-ray spectroscopies--with special attention paid to their strengths and weaknesses. Several specific examples of transition-metal nanocluster formation are then given, beginning with LaMer's classic sulfur sol system and including the Finke-Watzky mechanism of slow continuous nucleation A-->B followed by fast autocatalytic surface growth A+B-->2B. Finally, brief overviews of semiconductor nanoparticle preparations, solid-state nucleation studies-emanating from Avrami's work--and protein agglomeration mechanistic studies are also provided, as these processes are relevant, conceptually and in a general sense, to the field of transition-metal nanocluster nucleation and growth mechanisms.  相似文献   

5.
The reduction of [Ir(cod)Cl](2) (cod=1,5-cyclooctadiene) dissolved in 1-n-butyl-3-methyl tetrafluoroborate, hexafluorophosphate and trifluoromethane sulphonate ionic liquids in the presence of 1-decene by molecular hydrogen produces Ir(0) nanoparticles. The formation of these nanoparticles follows the two-step [A-->B, A+B-->2B (k(1),k(2))] autocatalytic mechanism. The same mean diameter values of around 2-3 nm were estimated from in situ TEM and SAXS analyses of the Ir(0) nanoparticles dispersed in the ionic liquids and by XRD of the isolated material. XPS and EXAFS analyses clearly show the interactions of the ionic liquid with the metal surface demonstrating the formation of an ionic liquid protective layer surrounding the iridium nanoparticles. SAXS analysis indicated the formation of an ionic liquid layer surrounding the metal particles with an extended molecular length of around 2.8-4.0 nm depending on the type of the anion.  相似文献   

6.
This study begins with the question of whether ionic liquids (ILs), such as 1-butyl-3-methylimidazolium hexafluorophosphate, [bmim][PF6], can be catalyst poisons for transition-metal catalysts rather than a preferred stabilizing media as typically assumed in the literature. The test case of acetone (propanone) hydrogenation is picked for two reasons: (i) acetone hydrogenation is important for its applications in heat pumps, H2 storage schemes, and fuel cells and for the commercial value of the resultant product, propan-2-ol, and (ii) two recent, independent studies have reported putative Ir(0)n nanocluster-catalyzed hydrogenations of acetone beginning in each case with the identical precursor, [{(COD)IrCl}2] (where COD=1,5-cyclooctadiene) (1). A close comparison of the results of those two literature studies and their related, but different, experimental conditions (vide infra) suggests the hypothesis that the IL is actually a catalyst poison. Indeed, the investigations herein (i) find that 1.0 equiv of added IL, [bmim][PF6], completely inhibits the formation of Ir(0)n nanoclusters under conditions 1 in Table 1 in the main text (namely, 3.6 mM precatalyst 1, 22 degrees C, and 2.76 bar H2) and (ii) demonstrate that 0.1 and 1.0 equivs of this same IL, [bmim][PF6], poisons 74 and 90%, respectively, of the acetone hydrogenation activity of premade, previously catalytically active nanoclusters. The above results in turn compelled a reinvestigation of the claim that Ir(0)n nanoclusters are the catalyst in what was reported as a colloidal suspension prepared under conditions 2 in Table 1 of the main text (namely, 52 mM precursor 1, 92 equiv of IL, 75 degrees C, and 4.05 bar of H2). We further (iii) find that the colloidal suspension prepared under conditions 2 is a mixture of unreacted precursor, 1, some nanoclusters, and isolable bulk metal, and we also (iv) find, somewhat surprisingly, in light of the IL-poisoning results found under conditions 1, that the Ir(0) catalyst prepared under conditions 2 is active, precisely as reported, for acetone hydrogenation. This, in turn, further demanded that we go on to (v) investigate the nature of the true catalyst under conditions 2, the results of which we are able to interpret only by the hypothesis that bulk metal is the dominant, true catalyst under conditions 2. Overall, the results provide strong evidence that ILs can be potent inhibitors of metal(0)/nanocluster catalysis, rather than the often-assumed superior solvent for nanocluster catalysis. The results also fortify our recent report that, under conditions where stoichiometrically high amounts of coordinating ligands are present (vs the amount of surface metal atoms), bulk-metal catalysts can actually be superior to nanocluster catalysts of the same metal, a seemingly heretical finding prior to our recent experimental evidence for this (Besson, C.; Finney, E. E.; Finke, R. G. J. Am. Chem. Soc. 2005, 127, 8179; Besson, C.; Finney, E. E.; Finke, R. G. Chem. Mater. 2005, 17, 4925).  相似文献   

7.
The complex [H(EtOH)2][{OsCl(eta4-COD)}2(mu-H)(mu-Cl)2] (1) has been prepared in high yield by treatment of OsCl3.3H2O (54% Os) with 1,5-cyclooctadiene in ethanol under reflux. Under air, it is unstable and undergoes oxidation by action of O2 to afford the neutral derivative {OsCl(eta4-COD)}2(mu-H)(mu-Cl)2 (2). The terminal chlorine ligands of the anion of 1 are activated toward nucleophilic substitution. Thus, reaction of the salt [NBu4][{OsCl(eta4-COD)}2(mu-H)(mu-Cl)2] (1a) with NaCp in toluene gives [NBu4][{Os(mu1-C5H5)(eta4-COD)}(mu-H)(mu-Cl)2{OsCl(eta4-COD)}] (3) as a result of the replacement of one of the terminal chlorine atoms by the cyclopentadienyl ligand. The CH2 group of the latter can be deprotonated by the bridging methoxy ligand of the iridium dimer [Ir(mu-OMe)(eta4-COD)]2. The reaction leads to the trinuclear derivative [NBu4][{(eta4-COD)Ir(mu5-C5H4-mu1)Os(eta4-COD)}(mu-H)(mu-Cl)2{OsCl(eta4-COD)}] (4) containing a bridging C5H4 ligand that is eta1-coordinated to an osmium atom of the dimeric unit and mu5-coordinated to the Ir(eta4-COD) moiety. Salt 1a also reacts with LiC[triple bond]CPh. In this case, the reaction produces the substitution of both terminal chlorine ligands to afford the bis(alkynyl) derivative [NBu4][{Os(C[triple bond]CPh)(eta4-COD)}2(mu-H)(mu-Cl)2] (5). Complexes 1, 2, 3, and 4 have been characterized by X-ray diffraction analysis. Although the separations between the osmium atoms are short, between 2.6696(4) and 2.8633(5) A, theoretical calculations indicate that only in 2 is there direct metal-metal interaction, as the bond order is 0.5.  相似文献   

8.
The synthesis and characterization of a previously unknown, rare organometallic-phosphate complex, {[Bu4N][(1,5-COD)Ir · HPO4]}n (1), is described. Characterization of 1 was accomplished by elemental analysis, electrospray mass spectrometry (ES-MS), and 1H and 13C NMR which established the symmetry of the product as at least C2 or Cs. The ES-MS reveals an interesting, Ir(I) to Ir(III) oxidative process with intense peaks displaying the 191Ir/193Ir isotopic distribution patterns expected for the fragments [(1,5-COD)IrIII(HPO4)2], [(C8H11)2(IrIII)2(PO4)(HPO4)(H2O)], and [(C8H11)2(IrIII)2(PO4)(HPO4)(H2O)2]. These fragments, in turn, provide evidence for a structure with two HPO42− groups attached to a single Ir, for example ring structures (of at least such C2 or Cs symmetry) such as {[Bu4N][(1,5-COD)Ir · HPO4]}2. Complex 1 is significant since it is known to be the preferred, compositionally precise precursor to the prototype example of a recently discovered class of novel, HPO42− and Bu4N+ stabilized nanoclusters, (Bu4N)2n2n+[Ir(0)n · (HPO4)n]2n. Such nanoclusters are being extended, via their analogous hydrogenphosphate-organometallic precursors (1,5-COD)M+ or 2+/HPO42− (M=Rh(I), Ru(II), Pt(II)) to their corresponding, catalytically active [M(0)n · (HPO4)n]2n nanoclusters.  相似文献   

9.
Following an introduction to the nanocluster stabilization literature and DLVO (Derjaugin-Landau-Verwey-Overbeek) theory of colloidal stability, the most common steric stabilizer of transition-metal nanoclusters, poly(vinylpyrrolidone) (PVP), has been examined for its efficacy in the formation, stabilization, and subsequent catalytic activity of prototype, test case Ir(0)n nanoclusters. First, the five criteria established previously for ranking nanocluster protectants for their nanocluster formation and stabilization ability were evaluated for 1 monomer equiv of 10000 average molecular weight (MWav) PVP in the absence, and then presence, of the traditionally weakly coordinating anion BF4- as well as the absence and presence of the strongly coordinating, superior anionic stabilizer P2W15Nb3O62(9-), all in propylene carbonate solvent. It is found that neither 1 equiv of BF4- in propylene carbonate nor 1 monomer equiv of (undried) PVP alone allows for isolable and redissolvable nanoclusters without bulk Ir(0)n metal formation. Careful predrying of the PVP, and by implication other polymers, is shown to be necessary for the formation and stabilization of the nanoclusters. Next, 40 monomer equiv of 10000 MWav PVP and 1 equiv of BF4- in propylene carbonate are shown to allow isolable, redissolvable nanoclusters. Control experiments reveal little difference on nanocluster stabilization by 3500 or 55000 (i.e., vs 10,000) MWav PVP, but yield interesting effects on nanocluster nucleation by the 3500 MWav PVP, as well as by the polymer poly(bis(ethoxy)phosphazene) (PBEP). Four other key polymers reported in the literature to be nanocluster stabilizers are tested by the five criteria method for their efficacy in the formation and stabilization of Ir0n nanoclusters (now in acetone due to the polymers' solubility) and in comparison to each other, specifically, poly(methyl methacrylate) (PMMA), poly(styrene) (PS), poly(methylhydrosilane) (PMHS), and PBEP. Only 40 monomer equiv dried PMMA allows isolable and redissolvable nanoclusters in acetone. Control/reference point experiments show that the electrostatic stabilizer P2W15Nb3O62(9-) is superior to each of the five polymeric stabilizers studied herein in both acetone and propylene carbonate, at least for the test case of Ir(0)n nanoclusters. Further controls show that 40 monomer equiv of PVP added to P2W15Nb3O(62)9--stabilized nanoclusters has no discernible effect on the five criteria other than to reduce by approximately 50% the nanocluster catalytic activity and total catalytic lifetime for cyclohexene hydrogenation. The main finding of this work is that DLVO theory as applied to nanocluster stabilization is fully supported; that is, surface-bound anions in high dielectric constant solvents provide superior stabilization. The importance of even traditionally weakly coordinating anions such as BF4- in nanocluster stabilization is a second, important finding of this work. The fact that HPO4(2-) has been shown to be a simple, cheap, commercially available, thermally robust, and 31P-NMR-handle-containing analogue of the more esoteric P2W15Nb3O62(9-) stabilizer is also discussed in the 14 total Conclusions from this first study ranking polymeric stabilizers of modern transition-metal nanoclusters.  相似文献   

10.
Acetone hydrogenation catalysis is important in applications such as heat pumps and fuel cells or in fulfilling the sizable demand for the product of selective acetone hydrogenation, 2-propanol. Reported herein is the discovery of a superior acetone hydrogenation catalyst--superior in terms of activity at low temperature, selectivity at complete conversion, and total catalyst lifetime. The new catalyst system consists of Ir(0)(n) nanoclusters plus HCl easily and reproducibly formed from commercially available [(1,5-COD)IrCl](2) under H(2). The resultant, room temperature, high activity, and highly selective (2/n)Ir(0)(n) plus 2HCl catalyst system hydrogenates acetone at 22 degrees C and 40 psig of H(2) pressure to 95% 2-propanol and the rest diisopropyl ether at 100% conversion with 16400 total catalytic turnovers and with an initial turnover frequency of 1.9 s(-1) at 22 degrees C. When molecular sieves are added, the catalyst system becomes even more selective and long-lived, providing the complete and selective conversion of acetone to 100% 2-propanol with 188000 total turnovers. Also reported are initial kinetic, D-labeling and other mechanistic studies, a summary section detailing the four main findings, the "green chemistry" aspects, and the current main drawback (a limited catalytic lifetime due to nanocluster precipitation) of the present invention. A review of the extensive literature of acetone hydrogenation is also tabulated as part of the Supporting Information.  相似文献   

11.
An improved, one-step synthesis of [RuII(1,5-COD)(CH3CN)4]2+ as the BF4 salt has been accomplished in 51% yield, an approximately 75% higher yield than the three-step literature synthesis of the corresponding PF6 salt. The improved synthesis consists of (i) grinding the insoluble [RuCl2(1,5-COD)]x precursor to increase the reaction rate and yield, (ii) treating the resultant [RuCl2(1,5-COD)]x with 2Ag+BF4 in refluxing acetonitrile with excess 1,5-COD present to inhibit 1,5-COD loss in the product and, most importantly, (iii) following the reaction directly by 1H-NMR spectrometry which revealed that the substitution reaction of the Ru(II), d6 precursor is, as expected, quite slow and requires ca. 120 h. The [Ru(1,5-COD)(CH3CN)4][BF4]2 product was characterized by 1H, 13C, and 19F-NMR, elemental analysis, and single-crystal X-ray crystallography. Problems in commercial Ru and F analyses are also addressed since this issue has been inadequately treated in the existing literature.  相似文献   

12.
The imidazolium salts: 1-mesityl-3-(3-trimethoxysilylpropyl)imidazolium iodide and 1-tert-butyl-3-(3-trimethoxysilylpropyl)imidazolium iodide, abbreviated as (tmpMes)HI (3a) and (tmp(t)Bu)HI (3b), respectively, have been synthesised. The palladium(ii) complexes (η(3)-C(3)H(5)) (tmpMes)PdCl (5a) and (η(3)-C(3)H(5))(tmp(t)Bu)PdCl (5b), rhodium(i) and iridium(i) complexes (η(4)-1,5-COD) (tmpMes)MCl, M = Rh (6a), Ir (7a) and (η(4)-1,5-COD)(tmp(t)Bu)MCl, where M = Rh (6b), Ir (7b), were synthesised by silver transmetallation reactions using the silver(i) complexes (tmpMes)AgI (4a) and (tmp(t)Bu)AgI (4b). The iridium complex 7b has been structurally characterised. The Pd(ii) and Rh(i) complexes have been immobilised by attachment to chemically modified MCM-41. The immobilised palladium(ii) materials have been tested as recyclable catalysts for Suzuki type C-C bond formation reactions in water and the immobilised rhodium(i) materials have been examined for their catalytic ability for the hydroformylation of 1-octene.  相似文献   

13.
Following a comprehensive look at the arene hydrogenation literature by soluble nanocluster catalysts, six key, unfulfilled goals in nanocluster arene hydrogenation catalysis are identified. To begin to address those six goals, well-characterized polyoxoanion- and tetrabutylammonium-stabilized Rh(0) nanoclusters have been synthesized by the reduction of the precisely defined precatalyst [Bu(4)N](5)Na(3)[(1,5-COD)Rh small middle dotP(2)W(15)Nb(3)O(62)] with H(2) in propylene carbonate solvent. These Rh(0) nanoclusters are characterized by their stoichiometry of formation, transmission electron microscopy, and the two rate constants which characterize their mechanism of formation; previous studies in our laboratories have provided additional characterization of polyoxoanion-stabilized Rh(0) nanoclusters. Propylene carbonate solutions of the Rh(0) nanoclusters catalyze the hydrogenation of anisole (methoxybenzene) under mild conditions (22-78 degrees C, 30-40 psig H(2)). Proton donors such as water or HBF(4) small middle dotEt(2)O are discovered to affect both nanocluster formation and nanocluster arene hydrogenation catalysis. Under identical conditions, the Rh(0) nanoclusters are 10-fold more active than a commercially available, oxide-supported 5% Rh/Al(2)O(3) catalyst of the same average metal-particle size. A series of lifetime experiments shows that the Rh(0) nanoclusters are capable of at least 2600 total turnovers (TTO), a lifetime significantly longer than the approximately 100 TTO often seen for nanocluster arene hydrogenation catalysts, and a lifetime slightly better than the prior record of 2000 TTO for a literature nanocluster system. The present polyoxoanion-stabilized Rh(0) nanoclusters also display a record, albeit modest, 30% selectivity for the partial hydrogenation of anisole to 1-methoxycyclohexene with an overall yield of up to 8% at higher temperatures. In comparison to the 5% Rh/Al(2)O(3) catalyst, the polyoxoanion-stabilized nanoclusters yield a 4.7-fold higher maximum yield of 1-methoxycyclohexene. Finally, the seven main findings of the present work are summarized, including how they address five of the aforementioned six main goals in nanocluster arene hydrogenation.  相似文献   

14.
To start, a brief introduction is provided on the importance of transition-metal nanoclusters, on the need to develop and then apply methods to rank the nanocluster formation and then stabilizing abilities of commonly employed anions, solvents, cations, and polymers, and on the somewhat confused literature of nanocluster stabilization. The fundamental importance of surface-adsorbed anions in transition-metal nanocluster stabilization is noted, the reason the present studies begin with a study of nanocluster-stabilizing anions. Next, five criteria, as well as the associated experimental methods, are developed to evaluate the efficacy of nanocluster stabilizing agents. The criteria are of fundamental significance in that they allow the separation of stabilizing agent effects on nanocluster formation from those on nanocluster stabilization. The results from applying the five criteria to four commonly employed anions lead to the first "anion series" of relative nanocluster-formation and stabilizing abilities, at least for the Ir(0) nanoclusters examined and by the following five criteria: [(P(2)W(15)Nb(3)O(61))(2)O](16-) (a Brphinsted-basic polyoxoanion) > C(6)H(5)O(7)(3-) (citrate trianion) > [-CH(2)-CH(CO(2))-](n)(n-) (polyacrylate) approximately Cl(-). In addition to the needed methods and the first anion series, six other (8 total) conclusions are reached, important insights in an area previously lacking hard information about which anions are the better choices for nanocluster formation and stabilization. The results are also of significance in establishing polyoxoanions, notably highly charged and basic polyoxoanions such as [(P(2)W(15)Nb(3)O(61))(2)O](16)(-), as the present "Gold Standards" among currently known nanocluster stabilizing anions, and according to the above five criteria. Such standards provide a reference point for future work aspiring to develop even better nanocluster stabilizing anions, solvents, cations, and polymers or their combinations.  相似文献   

15.
Oxo complex [(1,5-COD)4Pt4(mu3-O)2Cl2](BF4)2 (1) reacts readily with ethylene and norbornylene. The ethylene reaction yields acetaldehyde and a 1:1 mixture of (1,5-COD)Pt(Cl)(CH2CH3) (2) and [(1,5-COD)Pt4(eta3-CH2CHCH(CH3))](BF4) (3), while the norbornylene reaction yields a platinaoxetane complex, the first metallaoxetane to be obtained from the reaction of an oxo complex and an alkene.  相似文献   

16.
The isomerization of 1,5-cyclooctadiene (1,5-COD) to 1,4-COD and 1,3-COD catalyzed by (R-Cp)2TiCl2 (Cp = η5-C5H5)/R'MgX systems was studied. Cp2TiCl2/i-C3H7MgBr was found to have excellent catalytic activity for the isomerization of 1,5-COD to 1,3-COD at room temperature. The effects of solvent, Ti/Mg ratio, peroxide content in 1,5-COD, substituents on the cyclopentadienyl ligand of (R-Cp)2TiCl2, and alkyl groups in the Grignard reagents were examined and a titanium hydride addition-elimination mechanism was proposed.  相似文献   

17.
Oxazoline-thioether ligands 6-11 react with [Ir(η4-COD)Py2]PF6 (COD=C8H12=1,5-cyclooctadiene) to give [Ir(σ-η2-C8H12Py+)L] PF6 (L=oxazoline-thioether ligand) (12a-d) complexes resulted from the coordination of ligand to the metal and subsequent nucleophilic attack of pyridine to one of the double carbon bond of COD with concomitant iridium-carbon bond formation. When [Ir(η4-COD)2]BF4 was used as starting material, the reaction with ligands 7, 9 afforded the complexes [Ir(η4-COD)L]BF4. Application of these iridium complexes to the reduction of N-(α-methyl)benzylidenbenzylamine gave low or negligible enantioselectivity.  相似文献   

18.
Cleavage of the [Ir(η4-COD)Cl]2 dimer in the presence of the corresponding imidazolium salts and the strong base tBuO leads to the formation of Ir(I) derivatives of N-heterocyclic carbenes. When halide is replaced by NaCp, a mixture of [Ir(η4-COD)(NHCR)(η1-Cp)] and [Ir(η2-COD)(NHCR)(η5-Cp)] is obtained. The latter is favored for R = Cy, while the former predominates for R = Me. Conversely, [Ir(η4-COD)(NHCR)(η1-Ind)] is the only product of the reaction with NaInd, despite the R substituent. DFT/B3LYP calculations confirmed that the η1 coordination mode of the ring gives rise to the most stable structures, namely square planar complexes of 5d8 Ir(I). The energy of the 18 electron species containing η2-COD and η5-Ind or Cp is higher by 13 and 5 kcal mol−1, respectively. The fluxional behaviour of indenyl, detected by NMR in the solutions of [Ir(η4-COD)(NHCR)(η1-Ind)], is associated to the low energy of the η3-Ind species required in the conversion process, and is not easily observed in the cyclopentadienyl derivatives, where a similar intermediate is disfavored.  相似文献   

19.
The kinetics of formation of silver nanoparticles consisting of nearly 300 metal atoms is investigated, which were prepared by reduction of silver nitrate with hydrazine in ethylene glycol at 25°C without any stabilizer other than the glycol solvent. The resulting sigmoidal kinetic curves are analyzed by using the 1997 Finke–Watzky two-step mechanism of slow continuous nucleation with subsequent fast autocatalytic surface growth. The kinetics of homogeneous nucleation of metal nanoparticles was analyzed using the assumption about the stepwise adjunction of precursor and the quasi steady-state approximation. The equations were proposed to calculate the concentration of the formed metal nanoparticles and their mean size from the experimentally determined values of the Finke–Watzky rate constants. It is shown that a stepwise nucleation process can be described in the terms of the catalytically effective nucleus concept and that the number of atoms in the catalytically effective nucleus can be estimated.  相似文献   

20.
The theoretical infrared spectra of 1,3-cis-cis-cyclooctadiene (1,3-COD) and 1,5-cis-cis-cyclooctadiene (1,5-COD), were obtained by ab initio MO calculations at Hartree-Fock level. The results were compared with the available IR experimental spectra of 1,3- and 1,5-COD. The apparent agreement between theoretical and experimental data allows us to exploit two bands, found only in the case of the theoretical spectrum of 1,4-COD, as a tool for identifying 1,4-COD during its synthesis.  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