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.
Coordination of ambiphilic diphosphine–silane ligands [o‐(iPr2P)C6H4]2Si(R)F (R=F, Ph, Me) to AuCl affords pentacoordinate neutral silicon compounds in which the metal atom acts as a Lewis base. X‐ray diffraction analyses, NMR spectroscopy, and DFT calculations substantiate the presence of Au→Si interactions in these complexes, which result in trigonal‐bipyramidal geometries around silicon. The presence of a single electron‐withdrawing fluorine atom is sufficient to observe coordination of the silane as a σ‐acceptor ligand, provided it is positioned trans to gold. The nature of the second substituent at silicon (R=F, Ph, Me) has very little influence on the magnitude of the Au→Si interaction, in marked contrast to N→Si adducts. According to variable‐temperature and 2D EXSY NMR experiments, the apical/equatorial positions around silicon exchange in the slow regime of the NMR timescale. The two forms, with the fluorine atom in trans or cis position to gold, were characterized spectroscopically and the activation barrier for their interconversion was estimated. The bonding and relative stability of the two isomeric structures were assessed by DFT calculations.  相似文献   

2.
An Al/P‐based frustrated Lewis pair (FLP) reacted with PhMgCl by an unexpected transmetalation and formation of a phosphinylvinyl Grignard reagent. This compound is well suited for the transfer of the basic FLP component to other Lewis acidic metal atoms and allowed the generation of a Ga/P and an In/P2 FLP. The Ga FLP showed a behavior different to that of the corresponding Al FLP, the In FLP allowed the chelating coordination of an Au atom by Au−Cl bond activation.  相似文献   

3.
Borenium cations have been found to be valuable analogues of boranes as a result of their cationic character which imparts high electrophilicity. Herein, we report the synthesis, characterization, and reactivity of a new type of borenium cation employing a naphthyl bridge and a strong intramolecular P→B interaction. The cation reacts with H2 in the presence of PtBu3 (frustrated Lewis pair (FLP) approach) but also on its own. The mechanism of the reaction between the borenium cation and H2 in the absence of PtBu3 has been investigated using deuterium‐labeling experiments and DFT calculations. Both experiments and calculations imply the side‐on coordination of H2 to the B center, followed by heterolytic splitting and B? C bond cleavage. An uncommon syn 1,2‐carboboration has also been observed upon reaction of the borenium ion with 3‐hexyne.  相似文献   

4.
White‐light emitters have attracted considerable attention due to their importance in current and future technologies. By incorporating molecular fragments that independently emit in the blue, green/yellow, and red visible regions, specifically Cu‐NC, Au???Au interactions, and Cu‐SR2, respectively, into a single material, new white‐light‐emitting systems have been targeted. With this goal, three new CuI/thioether‐based coordination polymers containing bridging [Au(CN)2]? units have been synthesized and structurally characterized, and their photoluminescence properties (at room and low temperatures) have been delineated. Using this approach, white‐light emission (tunable from slightly yellow to slightly blue, depending on λex) is generated from Cu(Me2S)[Au(CN)2], a feature uncommon in such simple coordination compounds.  相似文献   

5.
Despite an absence of conventional porosity, the 1D coordination polymer [Ag4(O2C(CF2)2CF3)4(TMP)3] ( 1 ; TMP=tetramethylpyrazine) can absorb small alcohols from the vapour phase, which insert into Ag?O bonds to yield coordination polymers [Ag4(O2C(CF2)2CF3)4(TMP)3(ROH)2] ( 1‐ROH ; R=Me, Et, iPr). The reactions are reversible single‐crystal‐to‐single‐crystal transformations. Vapour‐solid equilibria have been examined by gas‐phase IR spectroscopy (K=5.68(9)×10?5 (MeOH), 9.5(3)×10?6 (EtOH), 6.14(5)×10?5 (iPrOH) at 295 K, 1 bar). Thermal analyses (TGA, DSC) have enabled quantitative comparison of two‐step reactions 1‐ROH → 1 → 2 , in which 2 is the 2D coordination polymer [Ag4(O2C(CF2)2CF3)4(TMP)2] formed by loss of TMP ligands exclusively from singly‐bridging sites. Four polymorphic forms of 1 ( 1‐ALT , 1‐AHT , 1‐BLT and 1‐BHT ; HT=high temperature, LT=low temperature) have been identified crystallographically. In situ powder X‐ray diffraction (PXRD) studies of the 1‐ROH → 1 → 2 transformations indicate the role of the HT polymorphs in these reactions. The structural relationship between polymorphs, involving changes in conformation of perfluoroalkyl chains and a change in orientation of entire polymers (A versus B forms), suggests a mechanism for the observed reactions and a pathway for guest transport within the fluorous layers. Consistent with this pathway, optical microscopy and AFM studies on single crystals of 1‐MeOH / 1‐AHT show that cracks parallel to the layers of interdigitated perfluoroalkyl chains develop during the MeOH release/uptake process.  相似文献   

6.
The oxoboryl complex trans‐[(Cy3P)2BrPt(B?O)] ( 2 ) reacts with the Group 13 Lewis acids EBr3 (E=Al, Ga, In) to form the 1:1 Lewis acid–base adducts trans‐[(Cy3P)2BrPt(B?OEBr3)] ( 6 – 8 ). This reactivity can be extended by using two equivalents of the respective Lewis acid EBr3 (E=Al, Ga) to form the 2:1 Lewis acid–base adducts trans‐[(Cy3P)2(Br3Al‐Br)Pt(B?OAlBr3)] ( 18 ) and trans‐[(Cy3P)2(Br3Ga‐Br)Pt(B?OGaBr3)] ( 15 ). Another reactivity pattern was demonstrated by coordinating two oxoboryl complexes 2 to InBr3, forming the 1:2 Lewis acid–base adduct trans‐[{(Cy3P)2BrPt(B?O)}2InBr3] ( 20 ). It was also possible to functionalize the B?O triple bond itself. Trimethylsilylisothiocyanate reacts with 2 in a 1,2‐dipolar addition to form the boryl complex trans‐[(Cy3P)2BrPt{B(NCS)(OSiMe3)}] ( 27 ).  相似文献   

7.
We present results from our investigations into correlating the styrene‐oxidation catalysis of atomically precise mixed‐ligand biicosahedral‐structure [Au25(PPh3)10(SC12H25)5Cl2]2+ (Au25bi) and thiol‐stabilized icosahedral core–shell‐structure [Au25(SCH2CH2Ph)18]? (Au25i) clusters with their electronic and atomic structure by using a combination of synchrotron radiation‐based X‐ray absorption fine‐structure spectroscopy (XAFS) and ultraviolet photoemission spectroscopy (UPS). Compared to bulk Au, XAFS revealed low Au–Au coordination, Au? Au bond contraction and higher d‐band vacancies in both the ligand‐stabilized Au clusters. The ligands were found not only to act as colloidal stabilizers, but also as d‐band electron acceptor for Au atoms. Au25bi clusters have a higher first‐shell Au coordination number than Au25i, whereas Au25bi and Au25i clusters have the same number of Au atoms. The UPS revealed a trend of narrower d‐band width, with apparent d‐band spin–orbit splitting and higher binding energy of d‐band center position for Au25bi and Au25i. We propose that the differences in their d‐band unoccupied state population are likely to be responsible for differences in their catalytic activity and selectivity. The findings reported herein help to understand the catalysis of atomically precise ligand‐stabilized metal clusters by correlating their atomic or electronic properties with catalytic activity.  相似文献   

8.
A series of lead(II) coordination polymers containing [N(CN)2]? (DCA) or [Au(CN)2]? bridging ligands and substituted terpyridine (terpy) ancillary ligands ([Pb(DCA)2] ( 1 ), [Pb(terpy)(DCA)2] ( 2 ), [Pb(terpy){Au(CN)2}2] ( 3 ), [Pb(4′‐chloro‐terpy){Au(CN)2}2] ( 4 ) and [Pb(4′‐bromo‐terpy)(μ‐OH2)0.5{Au(CN)2}2] ( 5 )) was spectroscopically examined by solid‐state 207Pb MAS NMR spectroscopy in order to characterise the structural and electronic changes associated with lead(II) lone‐pair activity. Two new compounds, 2 and [Pb(4′‐hydroxy‐terpy){Au(CN)2}2] ( 6 ), were prepared and structurally characterised. The series displays contrasting coordination environments, bridging ligands with differing basicities and structural and electronic effects that occur with various substitutions on the terpyridine ligand (for the [Au(CN)2]? polymers). 207Pb NMR spectra show an increase in both isotropic chemical shift and span (Ω) with increasing ligand basicity (from δiso=?3090 ppm and Ω=389 ppm for 1 (the least basic) to δiso=?1553 ppm and Ω=2238 ppm for 3 (the most basic)). The trends observed in 207Pb NMR data correlate with the coordination sphere anisotropy through comparison and quantification of the Pb? N bond lengths about the lead centre. Density functional theory calculations confirm that the more basic ligands result in greater p‐orbital character and show a strong correlation to the 207Pb NMR chemical shift parameters. Preliminary trends suggest that 207Pb NMR chemical shift anisotropy relates to the measured birefringence, given the established correlations with structure and lone‐pair activity.  相似文献   

9.
The reactivity of the free aluminylene [N]-Al ( 1 ) ([N]=1,8-bis(3,5-di-tert-butylphenyl)-3,6-di-tert-butylcarbazolyl) towards boron Lewis acids is investigated. A facile oxidative addition reaction of 1 with Ph2BOBPh2 furnishes an exceedingly scarce example of the free alumaborane [N]-Al(BPh2)(OBPh2) ( 2 ) with an Al−B electron-sharing bond. By contrast, complexation of 1 with B(C6F5)3 and HB(C6F5)2 gives rise to the corresponding Lewis adducts [N]-Al→B(C6F5)3 ( 3 ) and [N]-Al→BH(C6F5)2 ( 4 ), respectively, with an Al→B dative bond. Crystallization of 4 in Et2O produces the adduct [N]-Al(Et2O)→BH(C6F5)2 ( 5 ). Quantum chemical calculations are carried out to understand the formation of 2 as well as the bonding situation of 3 and 5 .  相似文献   

10.
Density functional theory method is used to examine a series of group III triazides X(N3)3 (X = B, Al, Ga). These compounds, except for the C3h planar B(N3)3 and Al(N3)3, are first reported here. C3h planar structures are the most energetically favored for all singlet X(N3)3 systems. Potential‐energy surfaces for unimolecular decompositions of the C3h and CS planar X(N3)3 species have been investigated. Results show that decomposition of B(N3)3 obeys sequential fashion and follows a four‐step mechanism: (1) B(N3)3 → NB(N3)2 + N2; (2) NB(N3)2 → cyc‐N2BN3 + N2; (3) cyc‐N2BN3 → trigonal‐BN3 + N2; (4) trigonal‐BN3 → linear‐NBNN. Decomposition of Al(N3)3 follows a two‐step mechanism: (1) Al(N3)3 → NAl(N3)2 + N2; (2) NAl(N3)2 → linear‐AlN3 + 2N2. The dissociation of Ga(N3)3 follows only one‐step mechanism: Ga(N3)3 → angular‐GaN3 + 3N2. These findings may be helpful in understanding the decomposition mechanisms of group III triazides as well as the possible mechanism for XN film generation. © 2014 Wiley Periodicals, Inc.  相似文献   

11.
Surface frustrated Lewis pairs (SFLPs) have been implicated in the gas‐phase heterogeneous (photo)catalytic hydrogenation of CO2 to CO and CH3OH by In2O3?x(OH)y. A key step in the reaction pathway is envisioned to be the heterolysis of H2 on a proximal Lewis acid–Lewis base pair, the SFLP, the chemistry of which is described as In???In‐OH + H2 → In‐OH2+???In‐H?. The product of the heterolysis, thought to be a protonated hydroxide Lewis base In‐OH2+ and a hydride coordinated Lewis acid In‐H?, can react with CO2 to form either CO or CH3OH. While the experimental and theoretical evidence is compelling for heterolysis of H2 on the SFLP, all conclusions derive from indirect proof, and direct observation remains lacking. Unexpectedly, we have discovered rhombohedral In2O3?x(OH)y can enable dissociation of H2 at room temperature, which allows its direct observation by several analytical techniques. The collected analytical results lean towards the heterolysis rather than the homolysis reaction pathway.  相似文献   

12.
Flexible, chelating bis(NHC) ligand 2 , able to accommodate both cis‐ and trans‐coordination modes, was used to synthesize ( 2 )Ni(η2‐cod), 3 . In reaction with GeCl2, it produced ( 2 )NiGeCl2, 4 , featuring a T‐shaped Ni0 and a pyramidal Ge center. Complex 4 could also be prepared from [( 2 )GeCl]Cl, 5 , and Ni(cod)2, in a reaction that formally involved Ni–Ge transmetalation, followed by coordination of the extruded GeCl2 moiety to Ni. A computational analysis showed that 4 possesses considerable multiconfigurational character and the Ni→Ge bond is formed through σ‐donation from the Ni 4s, 4p, and 3d orbitals to Ge. (NHC)2Ni(cod) complexes 9 and 10 , as well as (NHC)2GeCl2 derivative 11 , incorporating ligands that cannot accommodate a wide bite angle, failed to produce isolable Ni–Ge complexes. The isolation of ( 2 )Ni(η2‐Py), 12 , provides further evidence for the reluctance of the ( 2 )Ni0 fragment to act as a σ‐Lewis acid.  相似文献   

13.
Zeolites are highly important heterogeneous catalysts. Besides Brønsted SiOHAl acid sites, also framework AlFR Lewis acid sites are often found in their H‐forms. The formation of AlFR Lewis sites in zeolites is a key issue regarding their selectivity in acid‐catalyzed reactions. The local structures of AlFR Lewis sites in dehydrated zeolites and their precursors—“perturbed” AlFR atoms in hydrated zeolites—were studied by high‐resolution MAS NMR and FTIR spectroscopy and DFT/MM calculations. Perturbed framework Al atoms correspond to (SiO)3AlOH groups and are characterized by a broad 27Al NMR resonance (δi=59–62 ppm, CQ=5 MHz, and η=0.3–0.4) with a shoulder at 40 ppm in the 27Al MAS NMR spectrum. Dehydroxylation of (SiO)3AlOH occurs at mild temperatures and leads to the formation of AlFR Lewis sites tricoordinated to the zeolite framework. Al atoms of these (SiO)3Al Lewis sites exhibit an extremely broad 27Al NMR resonance (δi≈67 ppm, CQ≈20 MHz, and η≈0.1).  相似文献   

14.
The potential for coordination and H-transfer from Cp2MH2 (M=Zr, W) to gold(I) and gold(III) complexes was explored in a combined experimental and computational study. [(L)Au]+ cations react with Cp2WH2 giving [(L)Au(κ2-H2WCp2)]+ (L=IPr ( 1 ), cyclic (alkyl)(amino)carbene ( 2 ), PPh3 ( 3 ) and Dalphos-Me ( 4 ) [IPr=1,3-bis(diisopropylphenyl)imidazolylidene; Dalphos-Me=di(1-adamantyl)-2-(dimethylamino)phenyl-phosphine], while [Au(DMAP)2]+ (DMAP=p-dimethylaminopyridine) affords the C2-symmetric [Au(κ-H2WCp2)2]+ ( 5 ). The Dalphos complex 4 can be protonated to give the bicationic adduct 4 H, showing AuI⋅⋅⋅H+−N hydrogen bonding. The gold(III) Lewis acid [(C^N−CH)Au(C6F5)(OEt2)]+ binds Cp2WH2 to give an Au-H-W σ-complex. By contrast, the pincer species [(C^N^C)Au]+ adds Cp2WH2 by a purely dative W→Au bond, without Au⋅⋅⋅H interaction. The biphenylyl-based chelate [(C^C)Au]+ forms [(C^C)Au(μ-H)2WCp2]+, with two 2-electron-3-centre W−H⋅⋅⋅Au interactions and practically no Au−W donor acceptor contribution. In all these complexes, strong but polarized W−H bonds are maintained, without H-transfer to gold. On the other hand, the reactions of Cp2ZrH2 with gold complexes led in all cases to rapid H-transfer and formation of gold hydrides. Relativistic DFT calculations were used to rationalize the striking reactivity and bonding differences in these heterobimetallic hydride complexes along with an analysis of their characteristic NMR parameters and UV/Vis absorption properties.  相似文献   

15.
New auride Ca3Au3In was synthesized from the elements in a sealed tantalum tube in a high‐frequency furnace. Ca3Au3In was investigated by X‐ray powder and single crystal diffraction: ordered Ni4B3 type, Pnma, a = 1664.1(6), b = 457.3(2), c = 895.0(3) pm, wR2 = 0.0488, 1361 F2 values, and 44 variables. The three crystallographically independent boron positions of the Ni4B3 type are occupied by the gold atoms, while the four nickel sites are occupied by calcium and indium in an ordered manner. All gold atoms have trigonal prismatic coordination, i.e. Ca6 prisms for Au1 and Au2 and Ca4In2 prisms for Au3. While the Au3 atoms are isolated, we observe Au1–Au1 and Au2–Au2 zig‐zag chains at Au–Au distances of 292 and 284 pm. These slabs resemble the CrB type structure of CaAu. Consequently Ca3Au3In can be considered as a ternary auride. Together the Au2, Au3 and indium atoms build up a three‐dimensional [Au2In] polyanionic network (281–293 pm Au–In) in which the chains of Au1 centered trigonal prisms are embedded. The crystal chemical similarities with the structures of Ni4B3, CaAuIn, and CaAu are discussed.  相似文献   

16.
《Chemphyschem》2006,7(1):117-130
Ultra‐wideline 27Al NMR experiments are conducted on coordination compounds with 27Al nuclei possessing immense quadrupolar interactions that result from exceptionally nonspherical coordination environments. NMR spectra are acquired using a methodology involving frequency‐stepped, piecewise acquisition of NMR spectra with Hahn‐echo or quadrupolar Carr–Purcell Meiboom–Gill (QCPMG) pulse sequences, which is applicable to any half‐integer quadrupolar nucleus with extremely broad NMR powder patterns. Despite the large breadth of these central transition powder patterns, ranging from 250 to 700 kHz, the total experimental times are an order of magnitude less than previously reported experiments on analogous complexes with smaller quadrupolar interactions. The complexes examined feature three‐ or five‐coordinate aluminum sites: trismesitylaluminum (AlMes3), tris(bis(trimethylsilyl)amino)aluminum (Al(NTMS2)3), bis[dimethyl tetrahydrofurfuryloxide aluminum] ([Me2‐Al(μ‐OTHF)]2), and bis[diethyl tetrahydrofurfuryloxide aluminum] ([Et2‐Al(μ‐OTHF)]2). We report some of the largest 27Al quadrupolar coupling constants measured to date, with values of CQ(27Al) of 48.2(1), 36.3(1), 19.9(1), and 19.6(2) MHz for AlMes3 , Al(NTMS2)3 , [Me2‐Al(μ‐OTHF)]2 , and [Et2‐Al(μ‐OTHF)]2 , respectively. X‐ray crystallographic data and theoretical (Hartree–Fock and DFT) calculations of 27Al electric field gradient (EFG) tensors are utilized to examine the relationships between the quadrupolar interactions and molecular structure; in particular, the origin of the immense quadrupolar interaction in the three‐coordinate species is studied via analyses of molecular orbitals.  相似文献   

17.
Four new bis(benzimidazole)pyridine (BBP)‐containing compounds Zn(BBP)Cl[Au(CN)2], Mn(BBP)[Au(CN)2]2?H2O, Mn(BBP)Br2(MeOH) and Mn(BBP)Cl2(MeOH)?MeOH have been synthesized and structurally characterized and their birefringence values (Δn) determined. The structure of Zn(BBP)Cl[Au(CN)2] contains a hydrogen‐bonded dimer of Zn(BBP)Cl[Au(CN)2] units which propagate into a 1D chain through Au–Au interactions, although the crystals are of poor optical quality. The supramolecular structure of Mn(BBP)[Au(CN)2]2?H2O forms a 1D coordination polymer through chains of Mn(BBP)[Au(CN)2]2 units, each containing one bridging Au(CN)2 and one forming a 2D sheet through Au–Au interactions. The supramolecular structures of Mn(BBP)Br2(MeOH) and Mn(BBP)Cl2(MeOH)?MeOH are very similar, consisting of a complex hydrogen‐bonded network between NH imidazole, methanol and halide groups to align BBP building blocks. In the plane of the primary crystal growth direction, the birefringence values of the three Mn‐containing materials were Δn=0.08(1), 0.538(3) and 0.69(3), respectively. The latter two birefringence values are larger than in the related 2,2′;6′2′′‐terpyridine systems, placing them among the most birefringent solids reported. These compounds illustrate the utility of extending the π‐system of the building block and incorporating hydrogen‐bonding sites as design elements for highly birefringent materials and also illustrates the effect on the measurable birefringence of the crystal quality, growth direction and structural alignment of the anisotropic BBP building blocks.  相似文献   

18.
Models based on Au(111) face have been extensively used to describe self‐assembled monolayers, as well nanoparticles and nanoclusters. However, for very small clusters (<2 nm), the chemisorption of ligands leads to surface reconstruction, making necessary the use of a more reliable model that is able to simulate the main electronic and geometrical features of these small systems. In this work, a simple model to describe the geometries and the metal–ligand bonding in chalcogenate‐protected gold nanoclusters is proposed. Three different models with Aun+ and [XCH3]? (n=10, 15, 19, 22 and X=S, Se, Te) are used in this work. The obtained structures are in close agreement not only with the available crystallographic data, but also with much more expensive computational procedures, confirming that the proposed models are robust enough to describe the metal–ligand bonding. The results reveal that the Au–X distances are dependent on both the nature of the chalcogen and the coordination mode. The shortest Au–X distances are observed in the face‐centred cubic mode, indicating that the central gold atom seems to play a role in determining the adsorption strength. The proposed models show unambiguously chalcogen→cluster σ‐donation, as supported by energy decomposition analysis coupled with the natural orbitals for chemical valence and natural bond orbital analyses. In all cases, the metal–ligand interactions are characterised as being more covalent than electrostatic.  相似文献   

19.
The coordination chemistry of cyclic stannylene‐based intramolecular Lewis pairs is presented. The P→Sn adducts were treated with [Ni(COD)2] and [Pd(PCy3)2] (COD=1,5‐cyclooctadiene, PCy3=tricyclohexylphosphine). In the isolated coordination compounds the stannylene moiety acts either as an acceptor or a donor ligand. Examples of a dynamic switch between these two coordination modes of the P?Sn ligand are illustrated and the structures in the solid state together with heteronuclear NMR spectroscopic findings are discussed. In the case of a Ni0 complex, 119Sn Mössbauer spectroscopy of the uncoordinated and coordinated phosphastannirane ligand is presented.  相似文献   

20.
Two new oleanolic acid‐type triterpenoid saponins, raddeanosides R22 and R23 ( 1 and 2 , resp.), together with four known saponins were isolated from the rhizome of Anemone raddeana Regel. The structures of the new compounds were elucidated as oleanolic acid 3‐Oβ‐D ‐glucopyranosyl(1→2)[β‐D ‐glucopyranosyl(1→4)]‐α‐L ‐arabinopyranoside ( 1 ) and oleanolic acid 3‐Oα‐L ‐arabinopyranosyl(1→3)‐α‐L ‐rhamnopyranosyl(1→2)[β‐D ‐glucopyranosyl(1→4)]‐α‐L ‐arabinopyranoside ( 2 ). The four known compounds were identified as oleanolic acid 3‐Oα‐L ‐arabinopyranoside ( 3 ), oleanolic acid 3‐Oβ‐D ‐glucopyranosyl(1→4)‐α‐L ‐arabinopyranoside ( 4 ), hederasaponin B ( 5 ), and hederacholchiside E ( 6 ) on the basis of chemical and spectral evidences. Compound 4 is reported for the first time from the Anemone genus, while the other three known compounds have been already found in this plant.  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