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.
Corona[5]arenes, a novel type of macrocyclic compound that is composed of alternating heteroatoms and para ‐arylenes, were synthesized efficiently by two distinct methods. In a macrocycle‐to‐macrocycle transformation approach, S6‐corona[3]arene[3]tetrazine underwent sequential SNAr reactions with HS‐C6H4‐X‐C6H4‐SH (X=S, CH2, CMe2, SO2, and O) to produce the corresponding corona[3]arene[2]tetrazines. Different corona[3]arene[2]tetrazine compounds were also constructed in a straightforward manner by a one‐pot three‐component reaction of HS‐C6H4‐X‐C6H4‐SH (X=S, CH2, CMe2, SO2, and O) with diethyl 2,5‐dimercaptoterephthalate and 2 equiv of 3,6‐dichlorotetrazine under very mild conditions. All corona[5]arenes adopted 1,2,4‐alternate conformational structures in the crystalline state yielding similar nearly regular pentagonal cavities. Both the cavity size and the electronic property of the acquired macrocycles were fine‐tuned by the nature of the bridging element X.  相似文献   

2.
We report herein the synthesis, structure, and molecular recognition of S6‐ and (SO2)6‐corona[6](het)arenes, and demonstrate a unique and efficient strategy of regulating macrocyclic conformation and properties by adjusting the oxidation state of the heteroatom linkages. The one‐pot nucleophilic aromatic substitution reaction of 1,4‐benzenedithiol derivatives, biphenyl‐4,4′‐dithiol and 9,9‐dipropyl‐9H‐fluorene‐2,7‐dithiol with 3,6‐dichlorotetrazine afforded S6‐corona[3]arene[3]tetrazines. These compounds underwent inverse‐electron‐demand Diels–Alder reaction with enamines and norbornadiene to produce S6‐corona[3]arene[3]pyridazines. Facile oxidation of sulfide linkages yielded (SO2)6‐corona[3]arene[3]pyridazines. All corona[6](het)arenes adopted generally hexagonal macrocyclic ring structures; however, their electronic properties and conformation could be fine‐tuned by altering the oxidation state of the sulfur linkages. Whereas (SO2)6‐corona[3]arene[3]pyridazines were electron‐deficient, S6‐corona[3]arene[3]pyridazines acted as electron‐rich macrocyclic hosts that recognized various organic cations in both aqueous and organic solutions.  相似文献   

3.
Suzuki–Miyaura coupling reaction of BrC6H4-X-C6H4Br 1 (X=CH2, CO, N-Bu, O, S, SO, and SO2) with arylboronic acid 2 was investigated in the presence of tBu3PPd precatalyst and CsF/[18]crown-6 as a base to establish whether or not the Pd catalyst can undergo catalyst transfer on these functional groups. In the reaction of 1 (X=CH2, CO, N-Bu, O, and SO2) with 2 , aryl-disubstituted product 3 (Ar-C6H4-X-C6H4-Ar) was exclusively obtained, indicating that the Pd catalyst undergoes catalyst transfer on these functional groups. On the other hand, the reaction of 1 e (X=S) and 1 f (X=SO) with 2 afforded only aryl-monosubstituted product 4 (Ar-C6H4-X-C6H4-Br) and a mixture of 3 and 4 , respectively, indicating that S and SO interfere with intramolecular catalyst transfer. Furthermore, we found that Suzuki–Miyaura polycondensation of 1 (X=CH2, CO, N-Bu, O, and SO2) and phenylenediboronic acid 5 in the presence of tBu3PPd precatalyst afforded high-molecular-weight polymer even when excess 1 was used. The polymers obtained from 1 (X=CH2, N-Bu, and O) and 5 turned out to be cyclic.  相似文献   

4.
We report the efficient and scalable synthesis and molecular‐recognition properties of novel and water‐soluble S6‐corona[3]arene[3]pyridazines. The synthesis comprises a one‐pot nucleophilic aromatic substitution reaction between diesters of 2,5‐dimercaptoterephthalate and 3,6‐dichlorotetrazine followed by the inverse electron‐demand Diels–Alder reaction of the tetrazine moieties with an enamine and exhaustive saponification of esters. The resulting S6‐corona[3]arene[3]pyridazines, which adopt a 1,3,5‐alternate conformation in the crystalline state, are able to selectively form stable 1:1 complexes with dicationic guest species in water with association constants ranging from (1.10±0.06)×103 M ?1 to (1.18±0.06)×105 M ?1. The easy availability, large cavity size, strong and selective binding power render the water‐soluble S6‐corona[3]arene[3]pyridazines useful macrocyclic hosts in various disciplines of supramolecular chemistry.  相似文献   

5.
《Supramolecular Science》1996,3(4):189-205
Stereochemical problems and related functions of calix[4]arenes, calix[6]arenes and their chiral derivatives have been reviewed. In p-tert-butylcalix[4]arene (1H4) and its mono-, di-, tri-, and tetra-O-alkyl derivatives (1H3R, 1H2R2,1HR3, and 1R4, respectively), 23 different homologues can exist (including 1H4). We found that the OH group in the unmodified phenol unit is permeable through the calix[4]arene ring. Thus, several conformational isomers become equivalent after the ‘oxygen-through-the-annulus’ rotation of the OH group and the number of possible homologues is reduced to 13 (including 1H3). We report in this paper the syntheses of all of these possible conformational isomers using a protection-deprotection method with a benzyl group and metal template effects. On the other hand, all possible chiral isomers that can be derived from calix[4]arene by modification of the OH groups have been systematically classified. Molecular asymmetry can be generated not only by different substituents but also by conformational isomerism. The numbers of chiral isomers are 17 for tetra-O-substituted calix[4]arenes, 9 for tri-O-substituted calix[4]arenes, 3 for di-O-substituted calix[4]arenes, and 0 for mono-O-substituted calix[4]arenes. Chiral calix[4]arenes can also be designed by the introduction of a substituent into the m-position of a phenol unit or by the use of a dissymmetric ‘stapling reaction’ in proximal phenol units. In p-tert-butylcalix[6]arene, the conformational behaviour is totally different from that in p-tert-butylcalix[4]arene. A large degree of conformational freedom remains in the framework, and both ‘oxygen-through-the-annulus rotation’ and ‘para-substituent-through-the-annulus rotation’ can take place. However, when metal cations are bound to calix[6]aryl esters, the conformation is changed to a cone type. Bridging and capping are powerful methods to immobilize the conformation of calix[6]arenes. In addition, definitive evidence for ring immobilization was obtained from the absence of racemization in the chiral calix[6]arene. A successful example of chiral recognition for α-amino acid derivatives was achieved by using chiral homooxacalix[3]arene which has ‘pseudo C2 symmetry’. These examples indicate that calixarenes serve as rigid and conformationally diversiform platforms for the design of novel functional supramolecules.  相似文献   

6.
The inclusion of small neutral organic guests (C6H14, CH2Cl2, CH3OH) by calix[4]arene receptors was found by 1H NMR spectroscopy and microanalysis. The studied calix[4]arenes can form stable intramolecular complexes with solvent molecules which keep the stoichiometric composition without changing under conditions of the sublimation experiment. The saturated vapour pressures of calix[4]arenes and complexes of calix[4]arenes with solvent molecules were determinated for the first time by the Knudsen’s effusion method in the wide temperature range. The changing of standard thermodynamic parameters of complexation by transfer process from condensed state to vapour phase was estimated. It was shown that the large flexibility of the calixarene ligand structure corresponds to a strongly negative entropic contribution as well as negative enthalpy term to the Gibbs energy of formation of host–guest complexes in the gas phase.  相似文献   

7.
Procedures have been developed for the preparation of completely and partially adamantylated calix[n]arenes (n = 5, 6) by reaction of 3-R-substituted 1-hydroxyadamantanes (R = H, 4-MeC6H4, 4-MeSO2C6H4, 4-HO-3-HOCOC6H3, HOCOCH2) with p-H-calix[n]arenes (n = 5, 6) and 5,11,23,29-tetra-tert-butylcalix[6]arene in trifluoroacetic acid. Lower- and upper-rim modification of the prepared compounds has been studied. According to the 1H NMR data, adamantylcalix[6]arenes possessing carboxymethyl groups in the adamantane moieties are characterized by reduced conformational mobility.  相似文献   

8.
Reported here are the syntheses, conformational structures, electrochemical properties, and noncovalent anion binding of corona[5]arenes. A (3+2) fragment coupling reaction proceeded efficiently under mild reaction conditions to produce a number of novel heteroatom- and methylene-bridged corona[3]arene[2]tetrazine macrocycles. Selective oxidation of the sulfur atom between two phenylene rings afforded sulfoxide- and sulfone-linked corona[5]arenes in good yields. All corona[5]arenes synthesized adopted similar 1,2,4-alternate conformational structures, forming pentagonal cavities. The cavity sizes and the electronic properties such as redox potentials, were measured with CV and DPV, and were influenced by the different bridging units. As electron-deficient macrocycles, the acquired corona[3]arene[2]tetrazines served as highly selective hosts, forming complexes with the hydrogen-bonded dimer of dihydrogen phosphate through cooperative anion–π interactions.  相似文献   

9.
Metal Complexes of Biologically Important Ligands, CLVII [1] Halfsandwich Complexes of Isocyanoacetylamino acid esters and of Isocyanoacetyldi‐ and tripeptide esters (?Isocyanopeptides”?) N‐Isocyanoacetyl‐amino acid esters CNCH2C(O) NHCH(R)CO2CH3 (R = CH3, CH(CH3)2, CH2CH(CH3)2, CH2C6H5) and N‐isocyanoacetyl‐di‐ and tripeptide esters CNCH2C(O)NHCH(R1)C(O)NHCH(R2)CO2C2H5 and CNCH2C(O)NHCH(R1)C(O)NHCH (R2)C(O)NHCH(R3)CO2CH3 (R1 = R2 = R3 = CH2C6H5, R2 = H, CH2C6H5) are available by condensation of potassium isocyanoacetate with amino acid esters or peptide esters. These isocyanides form with chloro‐bridged complexes [(arene)M(Cl)(μ‐Cl)]2 (arene = Cp*, p‐cymene, M = Ir, Rh, Ru) in the presence of Ag[BF4] or Ag[CF3SO3] the cationic halfsandwich complexes [(arene)M(isocyanide)3]+X? (X = BF4, CF3SO3).  相似文献   

10.
Four new azocalix[4]arenes {5,11,17,23-tetrakis[(2-hydroxy-5-tert-butylphenylazo)]-25,26,27,28-tetrahydroxycalix[4]arene (1), 5,11,17,23-tetrakis[(2-hydroxy-5-nitro phenylazo)]-25,26,27,28-tetrahydroxycalix[4]arene (2), 5,11,17,23-tetrakis[(2-amino-5-carboxylphenylazo)]-25,26,27,28-tetrahydroxycalix[4]arene (3) and 5,11,17,23-tetrakis[(1-amino-2-hydroxy-4-sulfonicacidnapthylazo)]-25,26,27,28-tetrahydroxycalix[4]arene (4)} have been synthesized from p-tert-butylphenol, p-nitrophenol, p-aminobenzoic acid and 1-amino-2-hydroxy-4-sulphonic acid by diazo coupling reaction with p-aminocalix[4]arene. The resulting ligands (14) were treated with three transition metal salts (e.g., CuCl2·2H2O, NiCl2·6H2O or CoCl2·6H2O). Cu(II), Ni(II) and Co(II) complexes of the azocalix[4]arene derivatives were obtained and characterized by UV-vis, IR, 1H-NMR spectroscopic techniques and elemental analysis. All the complexes have a metal:ligand ratio of 2:1. The Cu(II) and Ni(II) complexes of azocalix[4]arenes are square-planar, while the Co(II) complexes of azocalix[4]arenes are octahedral with water molecules as axial ligands. The solvent extraction of various transition metal cations from the aqueous phase to the organic phase was carried out by using azocalix[4]arenes (14). It was found that, azocalix[4]arenes 1, 2 and 3 examined selectivity for transition metal cations such as Ag+, Hg+ and Hg2+. In addition, the thermal stability of metal:azocalix[4]arene complexes were also reported. Dedicated to Prof. Dr. Mustafa Yılmaz on the occasion of his 50th birthday  相似文献   

11.
Reported here are the syntheses, conformational structures, electrochemical properties, and noncovalent anion binding of corona[5]arenes. A (3+2) fragment coupling reaction proceeded efficiently under mild reaction conditions to produce a number of novel heteroatom‐ and methylene‐bridged corona[3]arene[2]tetrazine macrocycles. Selective oxidation of the sulfur atom between two phenylene rings afforded sulfoxide‐ and sulfone‐linked corona[5]arenes in good yields. All corona[5]arenes synthesized adopted similar 1,2,4‐alternate conformational structures, forming pentagonal cavities. The cavity sizes and the electronic properties such as redox potentials, were measured with CV and DPV, and were influenced by the different bridging units. As electron‐deficient macrocycles, the acquired corona[3]arene[2]tetrazines served as highly selective hosts, forming complexes with the hydrogen‐bonded dimer of dihydrogen phosphate through cooperative anion–π interactions.  相似文献   

12.
Reaction products have been isolated from SO2–L–H2O–О2 systems (L = ethylenediamine, N,N,N′,N′-tetramethylethylenediamine, piperazine, and morpholine) as onium salts [H3NCH2CH2NH3]SO4, [(CH3)2NHCH2CH2NH(CH3)2]SO4, [(CH3)2NHCH2CH2NH(CH3)2]S2O6 ? H2O, [C4H8N2H4]SO3 ? H2O, [C4H8N2H4]S2O6, [C4H8N2H4]SO4 ? H2O, [O(C2H4)2NH2]2SO4 ? H2O. The prepared compounds have been characterized by X-ray diffraction analysis, X-ray powder diffraction, IR and mass spectroscopy.  相似文献   

13.
Abstract

To gain information on CH-π aromatic interactions involved in the formation of host-guest adducts, the geometrical parameters which define the solid state structures of the complexes of calix[4]arenes in the cone conformation with guests having acid CH3 or CH2 groups have been studied. Most of the data have been obtained from the CH3CN and CH2Cl2 calix[4]arene complexes retrieved from the literature. To understand the effect of the acidity on these parameters, p-cyclohexylcalix[4]arene-biscrown-3 ? CH3CN, p-tert-butylthiacalix[4]arene ? CH3CN, p-tert-butylthiacalix[4]arene ? CH3NO2, 1,3-dipropoxy-p-tert-butylcalix[4]arene ? ClCH2CN and 1,3-dipropoxy-p-tert-butylcalix[4]arene ? CH2(CN)2 complexes were prepared, crystallised and investigated in the solid state. CH3X guests are bound preferentially by hosts having a C4 symmetry. The interaction is directional, but it is independent from the basicity of the host and acidity of the guest, indicating that classic hydrogen bond do not play a major role. On the contrary CH2XY guests find the best matching with hosts having a C2v symmetry, interacting specifically with two diametrical aromatic rings. These interactions are directional and show a correlation between the acidity of the guest and the CH-π aromatic distance, thus supporting a stronger contribution of “classic” hydrogen bond in these latter complexes. These results are in agreement with the hypothesis that CH-π aromatic interactions derive from the superimposition of different types of intermolecular forces, whose contribution depends on several factors as the nature of the interacting partners.  相似文献   

14.
Attempts to prepare previously unknown simple and very Lewis acidic [RZn]+[Al(ORF)4]? salts from ZnR2, AlR3, and HO?RF delivered the ion‐like RZn(Al(ORF)4) (R=Me, Et; RF=C(CF3)3) with a coordinated counterion, but never the ionic compound. Increasing the steric bulk in RZn+ to R=CH2CMe3, CH2SiMe3, or Cp*, thus attempting to induce ionization, failed and led only to reaction mixtures including anion decomposition. However, ionization of the ion‐like EtZn(Al(ORF)4) compound with arenes yielded the [EtZn(arene)2]+[Al(ORF)4]? salts with arene=toluene, mesitylene, or o‐difluorobenzene (o‐DFB)/toluene. In contrast to the ion‐like EtZn(η3‐C6H6)(CHB11Cl11), which co‐crystallizes with one benzene molecule, the less coordinating nature of the [Al(ORF)4]? anion allowed the ionization and preparation of the purely organometallic [EtZn(arene)2]+ cation. These stable materials have further applications as, for example, initiators of isobutene polymerization. DFT calculations to compare the Lewis acidities of the zinc cations to those of a large number of organometallic cations were performed on the basis of fluoride ion affinity. The complexation energetics of EtZn+ with arenes and THF was assessed and related to the experiments.  相似文献   

15.
To further our understanding of the properties of zinc-seamed pyrogallol[4]arene nanocapsules, we have investigated the energetic and geometric properties of the model complexes Zn(C2O2H3)2Y, Y = NH3, C5H5N, CH3OH, (CH3)2NCHO, or (CH3)2SO, with a zinc coordination sphere representative of that in the capsules. The effect of the choice of density functional, basis set, and zinc pseudopotential on the equilibrium structures and Y ligand bond dissociation enthalpies has been assessed. Among the ways in which the suitability of these models has been confirmed was the construction of polynuclear zinc complexes having 2, 4, 6, or 8 metal ions combined with C2O2H3 ?, C4O3H4 2?, and NH3 ligands, which indeed show that a closed ring is formed. The natural curvature of these complexes suggests that pentacoordination of Zn may be a key factor in seaming the pre-existing cone-shaped pyrogallol[4]arenes to form dimer capsules.  相似文献   

16.
O-Halogenosilyl-N,N-bis(trimethylsilyl)hydroxylamines – Synthesis, Crystal Structure, and Reactions The substitution of halogenosilanes on lithiated N,O-bis(trimethylsilyl)-hydroxylamine in the molar ratio of 1 : 1 occurs on the oxygen atom. The O-halogenosilyl-N,N-bis(trimethylsilyl)hydroxylamines were prepared: RSiF2ON · (SiMe3)2 (R = CMe3 1 , CHMe2 2 , CH2C6H5 3 , C6H2(CMe3)3 4 ), RR′SiFON(SiMe3)2 (R = CMe3, R′ = C6H5 5 ; R = Me, R′ = C6H5 6 ; R = C6H2Me3, R′ = C6H2Me3 7 ; R = CH2C6H5, R′ = CH2C6H5 8 ; R = CHMe2, R′ = CHMe2 9 ; R = CMe3, R′ = CMe3 10 ), RSiCl2ON(SiMe3)2 (R = CMe3 11 ; R = Cl 12 ). The reaction of fluorosilanes with lithiated N,O-bis(trimethylsilyl)hydroxylamine in the molar ratio of 1 : 2 leads to the formation of O,O′-fluorosilyl-bis[N,N-bis(trimethylsilyl)hydroxylamines]: RSiF[ON(SiMe3)2]2 (R = CMe3 13 ; R = C6H5 14 ). 13 could be prepared in the reaction of 1 with LiON(SiMe3)2. Lithiated dimethylketonoxime reacts with 1 to Me2C=NOSiRF–ON(SiMe3)2 [R = CMe3 ( 15 )]. The first crystal structure of a tris(silyl)hydroxylamine ( 4 ) is shown. The angle at the nitrogen prove a pyramidal geometry.  相似文献   

17.
The compounds [CH2(6-t-Bu-4-Me-C6H2O)2]PCl (1), (OCH2CMe2CH2O)-PCl (2), and [ClPN(t-Bu)]2 (3) have been utilized as precursors in the synthesis of (i) new pentacoordinate phosphorus compounds [e.g. CH2(6-t-Bu-4-Me-C6H2O)2 P(NRR′)(O2C6C14), CH2(6-t-Bu-4-Me-C6H2O)2PX[OC(O-i-Pr)N=N(C(O)O-i-Pr)],(ii) cyclic phosphates and their complexes [e.g. imidazolium+CH2(6-t-Bu-4-Me-C6H2O)2PO2 -.MeOH], (iii) new cycloaddition products [e.g. CH2(6-t-Bu-4-Me-C6H2O)2PC(CO2Me)C(CO2Me)C(O)N, (iv) macrocyclic compounds [e.g. [(t-BuN)P]2[-OCH2CMe2CH2O-]h2] and (v) phosphonates [e.g. (OCH2CMe2CH2O)P (O)CH2C(CN)=CHC5H4FeC5H5]. The synthetic and structural aspects of these new products are discussed.  相似文献   

18.
《中国化学》2018,36(7):630-634
O6‐Corona[3]arene[3]pyridazines were synthesized from the one‐pot macrocyclic condensation reaction of 3,6‐dichlorotetrazine with 1,4‐dihydroquinone derivatives followed by the inverse electron demand Diels‐Alder reaction of the tetrazine rings with a cyclopentanone‐derived enamine. Conversion of six ester groups within macrocycle into all sodium acetate moieties afforded a water soluble O6‐corona[3]arene[3]pyridazine. The coronary macrocycle host formed complexes selectively with organic ammoniums and dinitrile guests in a 1: 1 stoichiometric ratio in organic solvents with association constants ranging from (2.96 ± 0.10) × 101 to (2.53 ± 0.33) × 105 L·mol−1. Water soluble O6‐corona[3]arene[3]pyridazine was also able to complex strongly with organic ammoniums in water to give an association constant up to (2.67 ± 0.21) × 104 L·mol−1. The pseudo‐rotaxane and inclusion structures of the host‐guest complexes were revealed by the X‐ray crystallography.  相似文献   

19.
The following p-substituted N,N-bis-trimethlsilyl anilines p-X? C6H4? N[Si(CH3)3]2 are prepared by silylation of free amines: X = H, CH3, C2H5, CH3O, CH3CO, F, Cl, Br, J, CN, C6HS, (CH3)3SiO, and [(CH3)3Si]2N, and the isotopic derivatives C6H5? 15N[Si(CH3)3]2 and C6D5N[Si(CH3)3]2. The vibrational spectra are reported and assigned. The molecular symmetry of p-[(CH3)3Si]2N? C6H4? N[Si(CH3)3]2 is determined. The influence of the mass of the substituents X on the positions of the νsSiNSi vibrational frequencies is discussed.  相似文献   

20.
Treatment of [(1,5-C8H12)PtCl(X)] (X=Cl, CH3, CH2CMe3) with C2 chiral cyclopentane-1,2-diyl-bis(phosphanes) C5H8(PR2)2, either as racemic mixtures or as resolved enantiomers, afforded the chelate complexes [C5H8(PR2)2Pt(Cl)(X)] (X=Cl: R=Ph (1), N-pip (2), OPh (3); X=CH3: R=Ph (4), N-pip (5), OPh (6); X=CH2CMe3: R=Ph (7), N-pip (8), OPh (9); ‘N-pip’=N(CH2)5), (+)-[(1R,2R)-C5H8{P(OPh)2}2PtCl2] [(R,R)-3], (−)-[(1S,2S)-C5H8{P(OPh)2}2PtCl2] [(S,S)-3], (−)-[(1R,2R)-C5H8(PPh2)2Pt(Cl)(X)], and (+)-[(1S,2S)-C5H8(PPh2)2Pt(Cl)(X)] (X=CH3: (R,R)-4, (S,S)-4; X=CH2CMe3: (R,R)-7, (S,S)-7). Reacting 4, 6, and 7 with AgO3SCF3 led to triflate derivatives [C5H8(PR2)2Pt(X)(OSO2CF3)] [X=CH3: R=Ph (11), OPh (12); X=CH2CMe3: R=Ph (13)] with covalently bonded OSO2CF3 ligands. The unusual Pt2 complex [μ-Cl{C5H8(PPh2)2PtCH3}2]O3SCF3 (14) containing an unsupported single Pt---Cl---Pt bridge was also isolated. In the presence of SnCl2, complexes 1, 3, 4, 6, 7, and 9 are catalysts for the hydroformylation of styrene forming 2- and 3-phenylpropanal together with ethylbenzene. Except for 1, they also catalyze the consecutive hydrogenation of the primary propanals to alcohols. High regioselectivities towards 2-phenylpropanal (branched-to-normal ratios ≥91:9) were obtained in hydroformylations catalyzed by 3 and 4, for which the influence of varied CO/H2 partial pressures, catalyst-to-substrate ratios and different reaction temperatures and times on the outcome of the catalytic reaction was also studied. When tin-modified complexes (R,R)-3, (S,S)-3, and (S,S)-4 were used as optically active Pt(II) catalysts, an only low stereoselectivity for asymmetric hydroformylation (e.e.<18%) was observed. The Pt---Sn complexes [C5H8(PR2)2Pt(CH3)(SnCl3)] [R=Ph (15), OPh (17)], resulting from SnCl2 insertion into the Pt---Cl bonds of 4 or 6, undergo rapid degradation in solution, forming mixtures composed of [C5H8(PR2)2Pt(X)(Y)] with R=Ph or OPh and X/Y=Cl/SnCl3 (16, 18), Cl/Cl (1, 3), and SnCl3/SnCl3 (19, 20), respectively. In the presence of SnCl2, triflate complex 11 also becomes a catalyst for styrene hydroformylation and consecutive hydrogenation of the aldehydes to alcohols. The crystal structures of 11 complexes — 2, 5, 7, 8, 9, 10 (the previously prepared [C5H8{P(N-pip)2}2Pt(CH2CMe3)2]), 13, 14, 16, (R,R)-3, and (S,S)-3 — were determined by X-ray diffraction.  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