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 46 毫秒
1.
Phenolic 2-arylcoumarans 16 were used to examine the behaviors of β-5 subunits in lignin during tetramethylammonium hydroxide (TMAH) thermochemolysis. Products were monitored by gas chromatography/mass spectrometry. The process predominantly provided dimeric products with the opened hydrofuran ring. Substituent changes at the γ-position of ring A and at the 5-position of ring B had a large effect on the product compositions. 2-Arylcoumarans 1 and 6 with the γ-CH2OH substituent predominantly gave 2,3,3′,4′-tetramethoxystilbenes involving the elimination of the γ-CH2OH substituent, while 25 with the γ-CH3 substituent gave a mixture of 2,3,3′,4′-tetramethoxy-α-methylstilbenes and α-methoxy-α-(3′,4′-dimethoxyphenyl)-β-(2,3-dimethoxyphenyl)propanes. Substituent –CHCHCH3 on ring B remained unaffected. Substituents –CHCHCH2OH and –COOH on ring B produced the corresponding methyl ether and ester, respectively, by methylation. The –CHCHCHO substituent on ring B was converted to the –CHO substituent.  相似文献   

2.
A.P. Esteves 《Tetrahedron》2007,63(14):3006-3009
The controlled-potential reduction of [1-bromo-2-methoxy-2-(prop-2′-ynyloxy)ethyl]benzene (1a), 1-[2-bromo-2-phenyl-1-(prop-2′-ynyloxy)ethyl]-4-methoxybenzene (1b) and 2-bromo-3-(3′,4′-dimethoxyphenyl)-3-propargyloxypropanamide (1c) catalysed by (1,4,8,11-tetramethyl-1,4,8,11-tetraazacyclotetradecane)nickel(I), [Ni(tmc)]+, at a vitreous carbon cathode in DMF/Et4NBF4 leads to 2-methoxy-4-methylene-3-phenyl-tetrahydrofuran (2a), 2-(4′-methoxyphenyl)-4-methylene-3-phenyl-tetrahydrofuran (2b) and 2-(3′,4′-dimethoxyphenyl)-3-carbamoyl-4-methylenetetrahydrofuran (2c), respectively, in very high yields.  相似文献   

3.
Reaction of the Schiff base ligand derived from 4-pyridinecarboxaldehyde NC5H4C(H)N[2′,4′,6′-(CH3)C6H2], (1), with palladium(II) acetate in toluene at 60 °C for 24 h gave [Pd{NC5H4C(H)N[2′,4′,6′-(CH3)C6H2]}2(OCOCH3)2], (2), with two ligands coordinated through the pyridine nitrogen. Treatment of the Schiff base ligand derived from 4-pyridinecarboxaldehyde N-oxide, 4-(O)NC5H4C(H)N[2′,4′,6′-(CH3)C6H2], (4), with palladium(II) acetate in toluene at 75 °C gave the dinuclear acetato-bridged complex [Pd{4-(O)NC5H3C(H)N[2′,4′,6′-(CH3)C6H2]}(OCOCH3)]2, (5) with metallation of an aromatic phenyl carbon. Reaction of complex 5 with sodium chloride or lithium bromide gave the dinuclear halogen-bridged complexes [Pd{4-(O)NC5H3C(H)N[2′,4′,6′-(CH3)C6H2]}(Cl)]2, (6) and [Pd{4-(O)NC5H3C(H)N[2′,4′,6′-(CH3)C6H2]}(Br)]2, (7), after the metathesis reaction. Reaction of 6 and 7 with triphenylphosphine gave the mononuclear species [Pd{4-(O)NC5H3C(H)N[2′,4′,6′-(CH3)C6H2]}(Cl)(PPh3)], (8) and [Pd{4-(O)NC5H3C(H)N[2′,4′,6′-(CH3)C6H2]}-(Br)(PPh3)], (9), as air stable solids. Treatment of 6 and 7 with Ph2P(CH2)2PPh2 (dppe) in a complex/diphosphine 1:2 molar ratio gave the mononuclear complexes [Pd{4-(O)NC5H3C(H)N[2′,4′,6′-(CH3)C6H2]}(PPh2(CH2)2PPh2)][Cl], (10), and [Pd{4-(O)NC5H3C(H)N[2′,4′,6′-(CH3)C6H2]}(PPh2(CH2)2PPh2)][PF6], (11), with a chelating diphosphine. The molecular structure of complex 9 was determined by X-ray single crystal diffraction analysis.  相似文献   

4.
The rotational barriers between the configurational isomers of two structurally related push–pull 4-oxothiazolidines, differing in the number of exocyclic CC bonds, have been determined by dynamic 1H NMR spectroscopy. The equilibrium mixture of (5-ethoxycarbonylmethyl-4-oxothiazolidin-2-ylidene)-1-phenylethanone (1a) in CDCl3 at room temperature to 333 K consists of the E- and Z-isomers which are separated by an energy barrier ΔG# 98.5 kJ/mol (at 298 K). The variable-temperature 1H NMR data for the isomerization of ethyl (5-ethoxycarbonylmethylidene-4-oxothiazolidin-2-ylidene)ethanoate (2b) in DMSO-d6, possessing the two exocyclic CC bonds at the C(2)- and C(5)-positions, indicate that the rotational barrier ΔG# separating the (2E,5Z)-2b and (2Z,5Z)-2b isomers is 100.2 kJ/mol (at 298 K). In a polar solvent-dependent equilibrium the major (2Z,5Z)-form (>90%) is stabilized by the intermolecular resonance-assisted hydrogen bonding and strong 1,5-type S · · · O interactions within the SCCCO entity. The 13C NMR ΔδC(2)C(2′) values, ranging from 58 to 69 ppm in 1ad and 49-58 ppm in 2ad, correlate with the degree of the push-pull character of the exocyclic C(2)C(2′) bond, which increases with the electron withdrawing ability of the substituents at the vinylic C(2′) position in the following order: COPh COEt > CONHPh > CONHCH2CH2Ph. The decrease of the ΔδC(2)C(2′) values in 2ad has been discussed for the first time in terms of an estimation of the electron donor capacity of the S fragment on the polarization of the CC bonds.  相似文献   

5.
Treatment of the Schiff base ligands 4-(NC5H4)C6H4C(H)N[2′-(OH)C6H4] (a), 3,5-(N2C4H3)C6H4C(H)N[2′-(OH)-C6H4] (b) and 3,5-(N2C4H3)C6H4C(H) N[2′-(OH)-5′-tBuC6H3] (c) with palladium (II) acetate in toluene gave the poly-nuclear cyclometallated complexes [Pd{4-(NC5H4)C6H3C(H)N[2′-(O)C6H4]}]4 (1a), [Pd{3,5-(N2C4H3)C6H3C(H)N[2′-(O)-C6H4]}]4 (1b) and [Pd{3,5-(N2C4H3)C6H3C(H)N[2′-(O)-5′-tBuC6H3]}]4 (1c) respectively, as air stable solids, with the ligand acting as a terdentate [C,N,O] moiety after deprotonation of the –OH group. Reaction of the cyclometallated complexes with triphenylphosphine gave the mononuclear species [Pd{4-(NC5H4)C6H3C(H) N[2′-(O)C6H4]}(PPh3)], (2a), [Pd{3,5-(N2C4H3)C6H3C(H) N[2′-(O)C6H4]}(PPh3)], (2b) and [Pd{3,5-(N2C4H3)C6H3C(H)N[2′-(O)-5′-tBuC6H3)}(PPh3)], (2c) in which the polynuclear structure has been cleaved and the coordination of the ligand has not changed [C,N,O]. When the cyclometallated complexes 1b and 1c were treated with the diphosphines Ph2P(CH2)4PPh2 (dppb), Ph2PC5H4FeC5H4PPh2 (dppf) and Ph2P(CH2)2PPh2 (t-dppe) in a 1:2 molar ratio the dinuclear cyclometallated complexes [{Pd[3,5-(N2C4H3)C6H3C(H)N{2′-(O)C6H4}]}2(μ-Ph2P(CH2)4PPh2)], (3b), [{Pd[3,5-(N2C4H3)C6H3C(H) N{2′-(O)-5′-tBuC6H3}]}2(μ-Ph2P(CH2)4PPh2)], (3c), [{Pd[3,5-(N2C4H3)C6H3C(H)N{2′-(O)C6H4}]}2(μ-Ph2P(η5-C5H4)Fe(η5-C5H4)PPh2)], (4b), [{Pd[3,5-(N2C4H3)C6H3C(H) N{2′-(O)-5′-tBuC6H3}]}2(μ-Ph2P(η5C5H4)Fe(η5C5H4)P-Ph2)], (4c) and [{Pd[3,5-(N2C4H3)C6H3C(H)N{2′-(O)-5′-tBuC6H3}]}2(μ-Ph2P(CHCH)PPh2)], (5c) were obtained as air stable solids.  相似文献   

6.
Two approaches to the formation of ruthenium(II) complexes containing ligands with conjugated 2,2′:6′,2″-terpyridine (tpy), alkynyl and bithienyl units have been investigated. Bromination of 4′-(2,2′-bithien-5′-yl)-2,2′:6′,2″-terpyridine leads to 4′-(5-bromo-2,2’-bithien-5′-yl)-2,2′:6′,2″-terpyridine (1), the single crystal structure of which has been determined. The complexes [Ru(1)2][PF6]2 and [Ru(tpy)(1)][PF6]2 have been prepared and characterized. Sonogashira coupling of the bromo-substituent with (TIPS)CCH did not prove to be an efficient method of preparing the corresponding complexes with pendant alkynyl units. The reaction of 4′-ethynyl-2,2′:6’,2″-terpyridine with 5-bromo-2,2′-bithiophene under Sonogashira conditions yielded ligand 2, and the heteroleptic ruthenium(II) complex [Ru(2)(tpy)][PF6]2 has been prepared and characterized.  相似文献   

7.
Selective formation of (η3-siloxyallyl)tungsten complexes by reaction of hydrido(hydrosilylene)tungsten complexes with α,β-unsaturated carbonyl compounds was reported experimentally. The mechanisms have been investigated by employing the model reaction of [Cp(CO)2(H)WSi(H)–{C(SiH3)3}] (R), derived from the original experimental complex Cp′(CO)2(H)WSi(H)–[C(SiMe3)3] (1a, Cp′ = Cp*; 1b, Cp′ = η5-C5Me4Et), with methyl vinyl ketone, under the aid of the density functional calculations at the b3lyp level of theory. It is theoretically predicted that the route involving migration of the hydride to silicon to afford a 16e intermediate [Cp(CO)2W–SiH2–{C(SiH3)3}] is inaccessible (route 2), supporting the proposition by experiments. Another route, via [2 + 4] cycloaddition followed by directly Si–H reductive elimination, is theoretically predicted to be accessible (route 1). In route 1, two possible paths with different attacking directions of the oxygen of methyl vinyl ketone at Si (WSi) are put forward. The attack at the Si atom from the hydride (H1) side of the plane W–Si–H1 in R is found to be preferred kinetically. The regioselectivity for formation of (η3-siloxyallyl)tungsten complexes, where only the exo-anti isomer was obtained, is discussed based on the consideration of thermodynamics and kinetics.  相似文献   

8.
A comprehensive calculations were carried out to get a deep insight into the ground- and excited-state electronic structures and the spectroscopic properties for a series of [Pt(4-X–trpy)CCC6H4R]+ complexes (trpy = 2,2′,6′,2″-terpyridine; X = H, R = NO2 (1), Cl (2), C6H5 (3) and CH3 (4); R = Cl, X = CH3 (5) and C6H5 (6)). MP2 (second-order Møller–Plesset perturbation) and CIS (single-excitation configuration interaction) methods were employed to optimize the structures of 1–6 in the ground and excited states, respectively. The investigation showed that substituted phenylacetylide and trpy ligands only give rise to a small variation in geometrical structures but lead to a sizable difference in the electronic structures for 1–6 in the ground and excited states. The introduction of electron-rich groups into the phenylacetylide and/or terpyridyl ligands produces two different low-lying absorptions for 1 and 2–6, i.e., Pt(5d) → π*(trpy) metal-to-ligand charge transfer (MLCT) mixed with π → π*(CCPh) intraligand charge transfer (ILCT) for 1 and Pt(5d)/π(CCPh) → π*(trpy) charge transfer (MLCT and LLCT) for 26. Remarkable electronic resonance on the whole Pt–CCPh–NO2 moiety for 1 may be responsible for the difference. Solvatochromism calculation revealed that only LLCT/MLCT transitions showed the solvent dependence, consistent with the experimental observations.  相似文献   

9.
Studies on the catalytic reduction of nitrite on carbon electrodes modified with Co(II) tetra-2,3-pyridinoporphyrazine (CoTppa, 1), N,N′,N′′,N′′′-tetramethyltetra-2,3-pyridinoporphyrazine ([CoTm-2,3-tppa]4+, 2) and Co(II) N,N′,N′′,N′′′-tetramethyltetra-3,4-pyridinoporphyrazine ([CoTm-3,4-tppa]4+, 3) are reported. There is a close correspondence between the proximity of the methyl groups to the porphyrazine ring and the catalytic activity of the porphyrazine complexes. Bulk electrolysis gave ammonia and hydroxylamine as some of the products. The catalytic activity of the cationic complex, 3, towards the detection of low concentrations of nitrite (<10−9 M) in water containing sodium sulfate, was compared with the activities of the anionic cobalt(II) tetrasulfophthalocyanine ([CoTSPc]4−, 4) and the mixed [CoIITm-3,4-tppa]4+·[CoTSPc]4− (5) complexes. Complex 5 showed the best catalytic activity of the three in that large currents were obtained for very low concentrations of nitrite.  相似文献   

10.
[Bis[μ-[(2,3-butanedione dioximato)(2-)-O:O′]]tetrafluorodiborato(2-)-N,N′,N″,N]cobalt (COBF, 1) was used as a catalyst for the conversion of trans-1,2-dibromocyclohexane (DBCH) to cyclohexene and the photoelectrochemical cyclisation of 2-(4-bromobutyl)-2-cyclohexen-1-one (BBC) to trans-1-decalone 2 in a microemulsion. Voltammetry showed clear evidence of catalytic behaviour and bulk electrolysis showed larger turnover numbers for both reactions when compared with the same system using vitamin B12a as catalyst. For BBC, improved turnover may result from a relatively weak carbon–cobalt bond in the alkylcobalt intermediate of 1, and from better partition of 1 into the organic phase in which reactant BBC resides.  相似文献   

11.
The reaction dynamics of ground state boron atoms, B(2Pj), with acetylene, was reinvestigated and combined with novel electronic structure calculations. Our study suggests that the boron atom adds to the carbon–carbon triple bond of the acetylene molecule to yield initially a cyclic intermediate undergoing two successive hydrogen atom migrations to form ultimately an intermediate i3. The latter was found to decompose predominantly to the c-BC2H(X2A′) isomer plus atomic hydrogen via a tight exit transition state. To a minor amount, an isomerization of i3i4 prior to a hydrogen atom ejection forming the linear structure, HBCC(X1Σ+), has to be taken into account. Since the c-BC2H(X2A′) and HBCC(X1Σ+) isomers are separated by an isomerization barrier to ring closure of only 3 kJ mol−1, internally excited HBCC(X1Σ+) products can isomerize to the c-BC2H(X2A′) structure and vice versa.  相似文献   

12.
The reaction of 1-alkyl-2-{(o-thioalkyl)phenylazo}imidazoles (SRaaiNR) (2a/2b) with Ru(II) has synthesized [Ru(SRaaiNR)2](ClO4)2 (3a/3b) in 2-methoxyethanol. The reaction in methanol, however, has synthesized [Ru(SRaaiNR)(SRaaiNR)Cl](ClO4) (4a/4b). The solid phase reaction of SRaaiNR and RuCl3 on silica gel surface upon microwave irradiation has synthesized [Ru(SRaaiNR)(SaaiNR)](PF6) (5a/5b) [SRaaiNR represents tridentate N,N′,S-chelator; SRaaiNR is N,N′-bidentate chelator where S does not coordinate and SaaiNR refers N,N′,S-chelator where S refers to thiolato binding]. The structural characterization of [Ru(SEtaaiNEt)(SEtaaiNEt)Cl](ClO4) (4b) and [Ru(SEtaaiNEt)(SaaiNEt)](PF6) (5b) has been confirmed by single crystal X-ray diffraction study. The IR, UV–Vis, and 1H NMR spectral data also support the stereochemistry of the complexes. The complexes show metal oxidation, Ru(III)/Ru(II), and ligand reductions (azo/azo, azo/azo). The molecular orbital diagram has been drawn by density functional theory (DFT) calculation. Normal mode of analysis has been performed to correlate calculated and experimental frequencies of representative complexes. The electronic movement and assignment of electronic spectra have been carried out by TDDFT calculation both in gas and acetonitrile phase.  相似文献   

13.
The reaction of [Cp′Cr(CO)2(μ-SBu)]2 (1) (Cp′ = MeC5H4) with (PPh3)2Pt(PhCCPh) gives Cp′Cr(CO)2(μ-SBu)Pt(PPh3)2 (2) which could be regarded as a product of the substitution of acetylene ligand at platinum by a monomeric chromium–thiolate fragment. According to the X-ray diffraction analysis 2 contains single Cr–Pt (2.7538(15)) and Pt–S (2.294(2) Å) bonds while Cr–S bond (2.274(3) Å) is shortened in comparison with ordinary Cr–S bonds (2.4107(4)–2.4311(4) Å) in 1. The bonding between Cr–S fragment and platinum atom is similar to the olefine coordination in their platinum complexes.  相似文献   

14.
The interaction of amphiphilic cationic porphyrins, containing different patterns of meso-substitution by 4-(3-N,N,N-trimethylammoniumpropoxy)phenyl (A) and 4-(trifluoromethyl)phenyl (B) groups, with guanosine 5′-monophosphate (GMP) and calf thymus DNA have been studied by optical methods in phosphate buffer solution. The properties of these synthetic porphyrins were compared with those of representative standard of anionic 5,10,15,20-tetra(4-sulphonatophenyl)porphyrin (TPPS44−) and cationic 5,10,15,20-tetra(4-N,N,N-trimethylammonium phenyl)porphyrin (TMAP4+). Stable complexes with GMP were found for cationic porphyrins, except for monocationic AB3+. The binding constant (KGMP  104 M−1) follows the order: A3B3+  ABAB2+ > A44+  TMAP4+. Also, interaction with DNA was observed for all evaluated cationic porphyrins. For these related cationic porphyrins, the binding constant (KDNA  105 M−1) increases with the number of cationic charges. On the other hand, the photodynamic activity of porphyrins was analyzed in solution of GMP and DNA. Monocationic AB3+ is a less effective sensitizer to oxidize GMP in comparison with the other cationic porphyrins, in agreement with the lack of detected interaction with this nucleotide. The electrophoretic analysis of DNA indicates that photocleavage takes place when the samples are exposed to photoexcited tricationic and tetracationic porphyrins. In the presence of sodium azide the DNA decomposition was diminished. Also, reduction in the DNA photocleavage was observed under anoxic condition, indicating that oxygen is essential for DNA photocleavage sensitized by these cationic porphyrins. In addition, an increase in DNA degradation was not observed in deuteriated water. Therefore, an important contribution of type I photoreaction processes could be occurring in the DNA photodamage sensitized by these cationic porphyrins. These results provide a better understanding of the characteristics needed for sensitizers to produce efficient DNA photocleavage.  相似文献   

15.
Reactions of metal acetylide complexes M(CCAr)(PP)Cp′ (M = Fe, Ru; Ar = C6H5, C6H4Me-4; PP = (PPh3)2, dppe; Cp′ = Cp, Cp*; not all combinations), or the analogous vinylidene, with cyanogen bromide yield monobromovinylidene complexes [M{CC(Br)Ar}(PP)Cp′]+, isolated as PF6 salts. The trimethylsilyl-capped acetylides M(CCSiMe3)(PP)Cp′ react with cyanogen bromide to give [M(CCBr2)(PP)Cp′]+, the first examples of metal complexes containing a terminal dihalovinylidene ligand, which can be isolated as the BF4 salts. Molecular structures of representative mono- and di-bromovinylidene complexes are reported, together with those of Ru(CCSiMe3)(PPh3)2Cp and Ru(CCSiMe3)(dppe)Cp*.  相似文献   

16.
Substituted ethyl 1,2,3,4,4′,5′-hexahydrospiro[naphthalene-2,5′-pyrazole]-3′-carboxylates react with chlorine or N-bromosuccinimide to give spirocyclic substituted 3-halo-4,5-dihydro-3H-pyrazoles which lose nitrogen molecule on heating with formation of substituted spirocyclic 1-halocyclopropane-1-carboxylates. Heating of the title compounds with bromine in acetic acid results in opening of the spiro-fused six-membered ring to afford ethyl 4-aryl-5-[2-(2-carboxyphenyl)ethyl]pyrazole-3-carboxylates.__________Translated from Zhurnal Organicheskoi Khimii, Vol. 41, No. 7, 2005, pp. 1058–1063.Original Russian Text Copyright © 2005 by Molchanov, Korotkov, Kopf, Kostikov.  相似文献   

17.
The reaction of the heteroleptic Nd(III) iodide, [Nd(L′)(N″)(μ-I)] with the potassium salts of primary aryl amides [KN(H)Ar′] or [KN(H)Ar*] affords heteroleptic, structurally characterised, low-coordinate neodymium amides [Nd(L′)(N″)(N(H)Ar′)] and [Nd(L′)(N″)(N(H)Ar*)] cleanly (L′ = t-BuNCH2CH2[C{NC(SiMe3)CHNt-Bu}], N″ = N(SiMe3)2, Ar′ = 2,6-Dipp2C6H3, Dipp = 2,6-Pri2C6H3, Ar* = 2,6-(2,4,6-Pri3C6H2)2C6H3). The potassium terphenyl primary amide [KN(H)Ar*] is readily prepared and isolated, and structurally characterised. Treatment of these primary amide-containing compounds with alkali metal alkyl salts results in ligand exchange to give alkali metal primary amides and intractable heteroleptic Nd(III) alkyl compounds of the form [Nd(L′)(N″)(R)] (R = CH2SiMe3, Me). Attempted deprotonation of the Nd-bound primary amide in [Nd(L′)(N″)(N(H)Ar*)] with the less nucleophilic phosphazene superbase ButNP{NP(NMe2)3}3 resulted in indiscriminate deprotonations of peripheral ligand CH groups.  相似文献   

18.
The ground- and excited-state structures for a series of Os(II) diimine complexes [Os(NN)(CO)2I2] (NN = 2,2′-bipyridine (bpy) (1), 4,4′-di-tert-butyl-2,2′-bipyridine (dbubpy) (2), and 4,4′-dichlorine-2,2′-bipyridine (dclbpy) (3)) were optimized by the MP2 and CIS methods, respectively. The spectroscopic properties in dichloromethane solution were predicted at the time-dependent density functional theory (TD-DFT, B3LYP) level associated with the PCM solvent effect model. It was shown that the lowest-energy absorptions at 488, 469 and 539 nm for 13, respectively, were attributed to the admixture of the [dxy (Os) → π*(bpy)] (metal-to-ligand charge transfer, MLCT) and [p(I) → π*(bpy)] (interligand charge transfer, LLCT) transitions; their lowest-energy phosphorescent emissions at 610, 537 and 687 nm also have the 3MLCT/3LLCT transition characters. These results agree well with the experimental reports. The present investigation revealed that the variation of the substituents from H → t-Bu → Cl on the bipyridine ligand changes the emission energies by altering the energy level of HOMO and LUMO but does not change the transition natures.  相似文献   

19.
A DFT computational study is performed on different Cp2TiIV(L,L′-BID) complexes with L,L′-BID = dioxolene, dithiolene or diselenolene. A fragment analysis of the titanocene-ligand bonding in the DFT optimized geometries showed that out of plane folding for maximum Ti ← L π donation increases Cp2TiIV(O,O′-BID) (35°) < Cp2TiIV(S,S′-BID) (43–49°) < Cp2TiIV(Se,Se′-BID) (48–53°).  相似文献   

20.
Reaction of 2-(2′,6′-diethylphenylazo)-4-methylphenol (L2) with [Ir(PPh3)3Cl] afforded two organoiridium complexes 3 and 4 via C-H bond activation of an ethyl group in the arylazo fragment of the L2 ligand. In both the complexes the azo ligand binds to iridium as a dianionic tridentate C,N,O-donor. Two triphenylphosphines and a hydride (in the case of complex 3) or chloride (in the case of complex 4) are also coordinated to the metal center. A similar reaction of [Ir(PPh3)3Cl] with 2-(2′,6′-diisopropylphenylazo)-4-methylphenol (L3) yielded another organoiridium complex 5, where migration of one iso-propyl group from its original location (say, the 2′ position) to the corresponding third position (say, the 4′ position) took place through C-C bond activation. In this complex the modified azo ligand binds to iridium as a dianionic tridentate C,N,O-donor. Two triphenylphosphines and a hydride are also coordinated to the metal center. The structures of complexes 3 and 4 have been optimized through DFT calculations. The structure of complex 5 has been determined by X-ray crystallography. All the complexes show characteristic 1H NMR signals and intense transitions in the visible region. Cyclic voltammetry on all the complexes shows an oxidation within 0.66-1.10 V vs SCE, followed by a second oxidation within 1.15-1.33 V vs SCE and a reduction within −0.96 to −1.07 V vs SCE.  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