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 0 毫秒
1.
The acidity constants of the two-fold protonated acyclic 9-[2-(phosphonomethoxy)ethyl]-8-azaadenine, H2(9,8aPMEA)(+)(-), and its 8-isomer, 8-[2-(phosphonomethoxy)ethyl]-8-azaadenine, H2(8,8aPMEA)(+)(-), both abbreviated as H2(PA)(+)(-), as well as the stability constants of their M(H;PA)+ and M(PA) complexes with the metal ions M2+=Mg2+, Ca2+, Sr2+, Ba2+, Mn2+, Co2+, Ni2+, Cu2+, Zn2+ or Cd2+, have been determined by potentiometric pH titrations in aqueous solution at I=0.1 M (NaNO3) and 25 degrees C. Application of previously determined straight-line plots of log K(M)M(R-PO3) versus pK(H)H(R-PO3)for simple phosph(on)ate ligands, R-PO3(2-), where R represents a residue without an affinity for metal ions, proves that for all M(PA) complexes a larger stability is observed than is expected for a sole phosphonate coordination of the metal ion. This increased stability is attributed to the formation of five-membered chelates involving the ether oxygen present in the aliphatic residue (-CH2-O-CH2-PO3(2-)) of the ligands. The formation degrees of these chelates were calculated; they vary between about 13% for Ca(8,8aPMEA) and 71% for Cu(8,8aPMEA). The adenine residue has no influence on complex stability except in the Cu(9,8aPMEA) and Zn(9,8aPMEA) systems, where an additional stability increase attributable to the adenine residue is observed and equilibria between four different isomers exist. This means (1) an open isomer with a sole phosphonate coordination, M(PA)op, where PA(2-)=9,8aPMEA2-, (2) an isomer with a five-membered chelate involving the ether oxygen, M(PA)cl/O, (3) an isomer which contains five- and seven-membered chelates formed by coordination of the phosphonate group, the ether oxygen and the N3 site of the adenine residue, M(PA)cl/O/N3, and finally (4) a macrochelated isomer involving N7, M(PA)cl/N7. For Cu(9,8aPMEA) the formation degrees are 15, 30, 48 and 7% for Cu(PA)op, Cu(PA)cl/O, Cu(PA)cl/O/N3 and Cu(PA)cl/N7, respectively; this proves that the macrochelate involving N7 is a minority species. The situation for the Cu(PMEA) system, where PMEA2- represents the parent compound, i.e. the dianion of 9-[2-(phosphonomethoxy)ethyl]adenine, is quite similar. The relationship between the antiviral activity of acyclic nucleoside phosphonates and the structures of the various complexes is discussed and an explanation is offered why 9,8aPMEA is biologically active but 8,8aPMEA is not.  相似文献   

2.
The newly synthesised metal ion activated molecular receptor [Cd{1,4,7,10-tetrakis((R)-(−)-2-hydroxy-3-phenylpropyl)-1,4,7,10-tetraazacyclododecane}](ClO4)2 (hereafter [CdL](ClO4)2) acts as a molecular receptor for acetonitrile. The receptor was characterised by X-ray crystallography in its metal free form, as its Cd(II) complex with no included molecule, as its Cd(II) complex with an included acetonitrile molecule and, for comparative purposes, as its Zn(II) complex with a partially included acetonitrile molecule. These structural studies demonstrated that the Cd(II) complex is eight-coordinate, with the potential to form a well defined hydrophobic cavity that can contain one acetonitrile molecule through four hydrogen-bonds, whereas the Zn(II) complex is six-coordinate, with a less rigidly constituted binding cavity, such that when solvated by acetonitrile the solvating molecule is retained by only a single hydrogen-bond.  相似文献   

3.
New ligands containing a heterocyclic ring, L1 (1-anilino-2-(2-pyridyl)-naphth[1,2-d]imidazol-1-io-3-ide), L2 (2-phenyl-3-(2-pyridyl)-3,4-dihydro-naphtho[2,1-e][1,2,4]triazin-1-io-4-ide), and L3 (1-anilino-2-(2-quinolyl)-naphth[1,2-d]imidazol-1-io-3-ide), and their palladium (II) complexes have been prepared. Structures of the ligands and the complexes were determined by X-ray crystallography. The mononuclear square-planar complexes of [PdCl2(Ln)] (n = 1 (1), n = 2 (2) and n = 3 (3)) had didentate Ln (n = 1-3) ligands. The Ln (n = 1-3) ligands were stable and their absorption spectra did not change in dichloromethane and methanol. On the other hand, the absorption spectrum of [PdCl2(L2)] (2) in dichloromethane changed rapidly when methanol was added to the solution, and [PdCl(L4b)] (5) (L4b = N-[methoxy(2-pyridyl)methyl]-1-(phenylazo)-2-naphthylamide) was obtained from the concentrated reaction mixture. In this reaction, the dihydrotriazine ring of the didentate L2 ligand in complex 2 opened and the resulting tridentate L4b ligand coordinated to the Pd atom in complex 5. When an excess amount of (nBu)4NCl was added to complex 5 in dichloromethane, the absorption spectrum reverted to that of complex 2. Thus, the reversible ring opening and closure reactions of the coordinating dihydrotriazine ligand were observed. We also prepared [PdCl2(L5)] (9) (L5 = 1-(phenylazo)-N-[1-(2-pyridyl)ethylidene]-2-naphthylamine) and determined the structure. It is noted that neither the ring closure reaction nor the coordination of the azo nitrogen atom of the L5 ligand occurred in complex 9.  相似文献   

4.
Reaction of H2PtCl4 and K2PdCl4 with 2-hydroxyacetophenone N(4)-ethylthiosemicarbazone, H2Ap4Et, afforded [Pt(Ap4Et)(H2Ap4Et)] and [Pd(Ap4Et)(H2Ap4Et)]. Their crystal and molecular structures are reported and represent the first 1:2 thiosemicarbazone complexes with ligands having both different formal charge and denticity. The dianion, Ap4Et2−, coordinates in a planar conformation to palladium(II) or platinum(II) via the phenolato O, imine N and thiolato S atoms, while the neutral molecule exhibits monodentate coordination by the thione S atom. Intra-, intermolecular hydrogen bonds and C-H?π contacts lead to aggregation and a supramolecular assembly. Electronic, IR, and NMR spectral data, as well as electrochemical measurements, are included. The pKa values of the poorly water soluble H2Ap4Et were obtained spectrophotometrically in aqueous solutions of constant ionic strength.  相似文献   

5.
We have studied the binding interactions of biologically important carbohydrates (d-glucose, d-xylose and d-mannose) with the newly synthesized five-coordinate dinuclear copper(II) complex, [Cu2(hpnbpda)(μ-OAc)] (1) and zinc(II) complex, [Zn2(hpnbpda)(μ-OAc)] (2) [H3hpnbpda = N,N′-bis(2-pyridylmethyl)-2-hydroxy-1,3-propanediamine-N,N′-diacetic acid] in aqueous alkaline solution. The complexes 1 and 2 are fully characterized both in solid and solution using different analytical techniques. A geometrical optimization was made of the ligand H3hpnbpda and the complexes 1 and 2 by molecular mechanics (MM+) method in order to establish the stable conformations. All carbohydrates bind to the metal complexes in a 1:1 molar ratio. The binding events have been investigated by a combined approach of FTIR, UV–vis and 13C NMR spectroscopic techniques. UV–vis spectra indicate a significant blue shift of the absorption maximum of complex 1 during carbohydrate coordination highlighting the sugar binding ability of complex 1. The apparent binding constants of the substrate-bound copper(II) complexes have been determined from the UV–vis titration experiments. The binding ability and mode of binding of these sugar substrates with complex 2 are indicated by their characteristic coordination induced shift (CIS) values in 13C NMR spectra for carbon atoms C1, C2, and C3 of sugar substrates.  相似文献   

6.
Structures, chemical properties, and in vitro insulinomimetic activities of new vanadyl [oxovanadium(IV), VO(2+)] complexes with five tripodal ligands containing an imidazole functionality were examined. The ligands, N-(carboxymethyl)- N-(4-imidazolylmethyl)amino acids, contain glycine, ( S)- and ( R)-alanine, and ( S)- and ( R)-leucine residues. The molecular structures of the latter four alanine- and leucine-containing complexes were determined by X-ray analysis. The coordination geometry around each vanadium center was octahedral, where an imino nitrogen occupied the apical site and two carboxylate oxygens, an imidazole nitrogen, and a water molecule coordinated in the equatorial plane. The spectroscopic properties of the complexes were characterized by means of IR, electronic absorption, and CD spectra. Acid dissociation constants (p K(a)) and protonation sites of the ligands were determined by a combination of potentiometric titrations and (1)H NMR spectra. The potentiometric study demonstrated that stability constants (log beta) were not so different among the present complexes (14.0-14.9) and a species of molecular complex with a 1:1 metal:ligand ratio existed predominantly at physiological pH 7.4. EPR parameters indicated that the species at pH 7.4 had an octahedral structure similar to the complex in the solid state. On the other hand, an EPR study in phosphate buffer (pH 7.4) suggested that inorganic phosphate coordinated to the vanadium center instead of the imidazole group in the presence of excess phosphate ion. Cyclic voltammograms in the phosphate buffer showed chemically reversible oxidation waves, whereas irreversible oxidation waves were observed in non-coordinating HEPES buffer. Moreover, the oxidation potential of each complex in phosphate buffer was more positive than that in HEPES buffer. Partition coefficients of the present complexes in a n-octanol/saline system were very low, probably due to hydrophilicity of the imidazole group. The in vitro insulinomimetic activities were estimated on the basis of the ability of the complexes to inhibit epinephrine-stimulated free fatty acid release from isolated rat adipocytes. The achiral glycine-derivative complex exhibited the highest insulinomimetic activity, which was higher than that of VOSO(4) as a positive control. Putting our previous observations together, it was found that the vanadyl complexes with tetradentate amino acid derivatives having no alkyl side chain tend to have high in vitro insulinomimetic activity.  相似文献   

7.
The Zn(II)–cyclen–dipeptide ternary systems (where cyclen is abbreviated as L and dipeptide is glycylglycine (HL1) or glycyl‐(S)‐alanine (HL2)) were investigated by potentiometry applying both “out‐of‐cell” and direct titrations and by 1H NMR spectroscopy. Especially, the 1H NMR study was found to be very efficient to estimate speciation in the systems. The results obtained under full equilibria indicated two main species, [Zn(L)(HL1,2)]2+ and [Zn(L)(L1,2)]+, in both the systems. In the [Zn(L)(HL1,2)]2+ complex, presence of carbonyl‐carboxylate chelate was confirmed, and in the [Zn(L)(L1,2)]+ species, the peptide coordination is re‐organized to carbonyl‐amine chelate or only terminal amino group is coordinated. Equilibrium constants describing [Zn(L)]2+–dipeptide interaction are relatively low, log K = 3.4 for Gly‐Gly and 4.1 for Gly‐(S)‐Ala, respectively. Nevertheless, the values are slightly higher than stability constants for interaction of Zn(II) with the dipeptides (i.e. [Zn(L1,2)]+ species) where a chelate formation is expected. It indicates that interaction between Zn(II) ion in [Zn(L)]2+ and the dipeptides should be supported by some additional interactions. Potentiometry carried out under non‐equilibrum condition showed different species where these additional stabilizing forces play more important role. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

8.
Complex formation properties of a novel water soluble thiazolyloxime 2-(4-methylthiazol-2-yl)-2-(hydroxyimino)acetic acid (H3L1) with Cu2+ and Ni2+ were investigated in solution by potentiometrical and spectral (UV-Vis, EPR, NMR) methods. All Cu2+ and most of Ni2+ complex species detected in solution were found to have square-planar MN4 core with oxime and heterocyclic nitrogen atoms which was rationalized in terms of destabilizing effect of repulsive interaction between oxygen atom of carboxylic group and nitrogen atom of thiazole ring in N,O-coordinated ligand conformation. It has been found that stability of metal complexes in a series of oxime ligands is dependent upon basicity of nitrogen atom of oxime group. The thiazolyloxime forms less stable complexes with Cu2+ but stronger ones with Ni2+ ions when compared to parent 2-(hydroxyimino)propanoic acid. The lower stability obtained for Cu2+ complexes was elucidated in terms of negative inductive effect of the thiazole and nitrile substituents as well as an effect of intramolecular attractive interaction between thiazolyl sulfur and oxime oxygen atoms in thiazolyloxime. In the case of Ni2+ the complexes formed are square-planar and it is why thiazolyl ligand is more effective in metal ion binding than simple 2-(hydroxyimino)propanoic acid forming only octahedral species. The solid state structure of the Co3+ complex K3[Co(HL1)3]·5.5H2O (1) was studied by X-ray analysis. The thiazolyloxime ligand is coordinated to Co3+ via oxime nitrogen and carboxylate oxygen atoms forming five-membered chelate rings.  相似文献   

9.
Abstract

Tricyclic (T) analogues of acyclovir (ACV, 1) and ganciclovir (GCV, 2) carrying the 3,9-dihydro-9-oxo-5H-imidazo[1,2-a]purine system [i.e., 6-(4-BrPh)TACV, 5 and 6-(4-BrPh)TGCV, 6] were transformed into 6-[(4′-R2)-4-biphenylyl] derivatives of TACV (79) and TGCV (1012) by Suzuki cross coupling with 4-substituted phenylboronic acids. Compound 11 (R2 = CH2OH) showed a high (~1000) selectivity index against herpes simplex virus type 1 (HSV-1) together with advantageous fluorescence properties (emission in visible region, little overlap with absorption and moderate intensity).  相似文献   

10.
(-)-9-Fluorocytisine, (-)-9-methylcytisine and (-)-9-trifluoromethylcytisine were synthesized from the natural product (-)-cytisine. 9-Methyl and 9-trifluoromethyl cytisines display a remarkable affinity at the α(4)β(2) nicotinic receptor subtype (0.2 nM) with a high selectivity versus the α(7) nAChR subtype. Comparison of the affinity values suggests that the size of the substituent at the 9 position of (-)-cytisine seems more important than electronic factors for efficient binding and selectivity at α(4)β(2) nAChRs.  相似文献   

11.
The tetragonal-pyramidal VO2+ complexes [VO{(RSC-S)N-NX}2] (1-6) were synthesised by the reactions of VO(OCHMe2)3 with the dithiocarbazate ligands RSC(S)-NH-NX, where X = cyclo-pentyl, cyclo-hexyl or 4-Me2N-C6H4-CH, and R = CH3 or CH2C6H5. The compounds were characterised by elemental analysis, IR- and mass spectrometries, and in cases of compounds 1, 3, 4 and 5, by X-ray diffraction. The chiral compound 4 (X = cyclo-hexyl, R = CH2C6H5) crystallises in the C configuration. In compound 5, the VO moiety is disordered (83.3:16.7%) with respect to the plane spanned by the four equatorial ligand functions.  相似文献   

12.
The stoichiometries and the affinity toward simple and paired metal ions of synthetic amyloid-β(1-40) peptide (Aβ1-40) were investigated by electrospray ion trap mass spectrometry (ESI-MS), circular dichroism (CD), and atomic force microscopy (AFM). The results lead to the working hypothesis that pH-dependent metal binding to Aβ1-40 may induce conformational changes, which affect the affinity toward other metals. A significant copper and zinc binding to Aβ1-40 peptide at pH 5.5 was found, whereas nickel ions commonly bind to each molecule of β-amyloid peptide. Some complexes of Aβ1-40 with more than one nickel ion were identified by ESI-MS. In addition, nickel ions proved to enhance Aβ oligomerization. On increasing pH, up to 12 ions of zinc may bind to a single Aβ molecule. Under the same pH and concentration conditions, the binding pattern of the independent copper and silver ions to Aβ1-40 was different from that of the equimolecular mixture of the two metal ions. One might assume that some conformational changes due to water loss altered the capacity of Aβ peptide to bind certain heavy metal ions. As a consequence, copper–silver interaction with the binding process to Aβ1-40 became highly complex. A competition between silver and nickel ions for Aβ1-40 binding sites at high pH was also observed. New strategies were proposed to identify the characteristic signals for some important metal ion–peptide complexes in the spectra recorded at high pH or high concentrations of metal ions. To explain the formation of such a large number of high metal ion–Aβ complexes, we took into consideration the participation of both histidine residues and free amino groups as well as carboxylate ones in the binding process. Finally, CD and AFM studies supported the mass spectrometric data.  相似文献   

13.
As major effector cells of the innate immune response, macrophages must adeptly migrate from blood to infected tissues. Endothelial transmigration is accomplished by matrix metalloproteinase (MMP)-induced degradation of basement membrane and extracellular matrix components. The classical activation of macrophages with LPS and IFN-γ causes enhanced microtubule (MT) stabilization and secretion of MMPs. Macrophages up-regulate MMP-9 expression and secretion upon immunological challenge and require its activity for migration during the inflammatory response. However, the dynamics of MMP-9 production and intracellular distribution as well as the mechanisms responsible for its trafficking are unknown. Using immunofluorescent imaging, we localized intracellular MMP-9 to small Golgi-derived cytoplasmic vesicles that contained calreticulin and protein-disulfide isomerase in activated RAW 264.7 macrophages. We demonstrated vesicular organelles of MMP-9 aligned along stable subsets of MTs and showed that selective modulation of MT dynamics contributes to the enhanced trafficking of MMP-9 extracellularly. We found a Rab3D-dependent association of MMP-9 vesicles with the molecular motor kinesin, whose association with the MT network was greatly enhanced after macrophage activation. Finally, we implicated kinesin 5B and 3B isoforms in the effective trafficking of MMP-9 extracellularly.  相似文献   

14.
 The stability constants of the 1 : 1 complexes formed between Mg2+ and the anions of the N1, N3, and N7 deaza derivatives of 9-[2-(phosphonomethoxy)ethyl]adenine (PA2–), i.e., of Mg(H;PA)+ and Mg(PA), were determined by potentiometric pH titration in aqueous solution (25  °C; I=0.1 M, NaNO3) and compared with previous results [Sigel H, et al. (1992) Helv Chim Acta 75 : 2634–2656], obtained under the same conditions, for the corresponding complexes of 9-[2-(phosphonomethoxy)ethyl]adenine (PMEA2–) and (phosphonomethoxy)ethane (PME2–). Based on the analysis of a microconstant scheme it is concluded that in the monoprotonated complexes, Mg(H;PA)+, Mg2+ is coordinated to a significant part at the nucleobase, H+ being at the phosphonate group. By making use of log K Mg Mg(R-PO3) versus pK H H(R-PO3) straight-line plots (also obtained previously; see above) for simple phosphonates and phosphate monoesters, it is shown that all the Mg(PA) complexes, including those with PMEA2– and PME2–, are more stable than expected on the basis of the basicity of the ―PO2– 3 group. This proves that, to some extent, five-membered chelates, Mg(PA)cl/O, involving the ether oxygen of the ―CH2―O―CH2―PO2– 3 chain are formed; their formation degree amounts to about 30–40% in equilibrium with the isomer having only a phosphonate-Mg2+ coordination. In the case of Mg(1-deaza-PMEA), probably a further isomer occurs in which also N3 of the nucleobase participates. The different properties between the Mg(PA) species and the Mg(AMP) complex are discussed. Received: 26 January 1998 / Accepted: 19 May 1998  相似文献   

15.
The synthesis and characterization of mononuclear copper complex of (3-oxo-2,3-dihydro-benzo[1,4]oxazin-4-yl)-acetic acid (1) and a tetranuclear copper complex of (2-carboxymethoxy-phenylamino)acetic acid (2) is reported. The sodium salt 1 on reaction with copper(II) chloride hexahydrate followed by treatment with pyridine gave a mononuclear copper complex; whereas, a tetra-nuclear complex in the case of reaction of 2 with copper(II) chloride hexahydrate and 2,2′-bipyridine was obtained. In tetra-nuclear copper(II) complex the NH group co-ordinates to copper and cluster has five co-ordination around copper(II).  相似文献   

16.
The synthesis and properties of the copper(II) complex with dimethylaminomethylphosphine oxide as a ligand will be presented. The complex, with the formula [Cu(NO3)2(POC3H10N)2] 1, has been characterized by elemental analysis, IR spectroscopy, SQUID and X-ray measurements. The X-ray structure was determined for the complex 1 and for the ligand dimethylaminomethylphosphine oxide (DMAO) 2. The single crystal X-ray structure of 1 shows that in the crystal the copper ions form distorted octahedral environment, consisting of two oxygen and two nitrogen atoms from the DMAO ligand. Additionally, one oxygen atom of each anion is semi-coordinated to the copper ion. The solid state magnetic measurements show that the complex 1 is paramagnetic with weak antiferromagnetic interactions in low temperatures.  相似文献   

17.
Optically active enamines of 2-(2′-pyrido)acetophenone or 2-(2′-quinolino)acetophenone with (R)-1-phenylethylamine, (R)-1-(1-naphthyl)ethylamine, (R)-cyclohexylethylamine, and (R)-phenylglycinol were prepared and their copper(I) complexes used in the enantioselective cyclopropanation of styrene with ethyl- and menthyldiazoacetate. Enantioselectivities of up to 42% enantiomeric excess were obtained for cis/trans 2-phenylcyclopropan-1-carboxylic acid ethyl esters, as determined by gas-liquid chromatography (GLC) on chiral chromatographic columns. © 1995 Wiley-Liss, Inc.  相似文献   

18.
The template reaction between salicylaldehyde S-methyl-isothiosemicarbazone and 2-formylpyridine in presence of nickel(II) or copper(II) salts yields two new coordination compounds with general formula [NiL1]2(1) and [CuL2]2(2) (L1 = the dianionic (N1-salicylidene)(N4-(hydroxy(pyridin-2-yl)methyl) S-methyl-isothiosemicarbazide) ligand and L2 = the doubly deprotonated (N1-salicylidene)(N4-(picolinoyl) S-methyl-isothiosemicarbazide) ligand). In the complex 1, the formed L1 ligand appears as result of an addition reaction of the precursors, while for 2 a redox mechanism is implicated in the formation of L2. Despite the fact that the initial organic precursors are the same, the resulting ligands obtained in the template reaction are different. In 1, the Ni(II) metal ion adopts a square-planar geometry and the [NiL1] units are forming dimerized chains through weak Ni···Ni interactions (3.336 and 3.632 Å). In 2, the Cu(II) metal ions adopt a square-pyramidal geometry and form dinuclear species through weak Cu···O (phenoxo) interactions. The magnetic susceptibility measurements of the complexes reveal the diamagnetic nature of 1 as expected for a square planar Ni(II) complex and a paramagnetic behavior for 2 with weak intra-dimer antiferromagnetic interaction (J/kB = −2.1(1) K).  相似文献   

19.
The synthesis and spectroscopic (IR, (1)H and (13)C NMR) characterization of new complexes of Pt(II), Pd(II), Cu(II), and Hg(II) with the Schiff base ligand MeCONHCH(2)CH(2)N=CHPy (L) (Py=pyridine) are reported, together with studies on the cytotoxicities of these complexes, L and [ReBr(CO)(3)(L)] against human leukemia (MOLT-4), breast cancer (MCF-7) and Chang Liver (non-cancerous) cells. The crystal structures of [Pt(L)Cl(2)] (2), [Cu(L)Cl(2)] (4) and [Hg(L)Cl(2)](2) (5) are also reported. Of the complexes studied, [Cu(L)Cl(2)] (4) was identified as the most cytotoxic active derivative against cells of neoplastic origin (MOLT-4, and MCF-7), while having low toxicity on cells of benign origin (Chang Liver).  相似文献   

20.
Four novel imidazolium salts, precursors to N-heterocyclic carbene (NHC) ligands, with 2,5-dimethoxybenzyl or 2,5-dihydroxybenzyl (i.e., p-hydroquinone) substituents have been prepared. The crystal structure of the hydroquinone-substituted imidazolium salt H3L3Br reveals Br?H-O bridged chiral chains of alternating [H3L3]+ cations and Br counter-ions parallel to the x-axis. Palladium(II) complexes were accessible from reactions of the dimethoxyphenyl-substituted imidazolium precursors with palladium(II) acetate, but not from reactions of imidazolium cations with hydroquinonyl substituents. The crystal structure of the bis(dimethoxybenzyl)-substituted bis(NHC)Pd complex, cis-[PdBr2(L2)] (2), is described. Puckering of the bis(NHC) ligand leads to a cleft in which an included molecule of dimethylformamide is situated. The cleft is closed by one of the dimethoxybenzyl groups which π-stacks with the dimethylformamide; the other dimethoxybenzyl group points away from the cleft and Pd(II) centre. Reaction of complex 2 with BBr3 afforded the targeted bis(hydroquinone)-substituted bis(NHC)Pd(II) complex 3 (97% yield) which, in turn, was oxidised by 2,3-dichloro-5,6-dicyano-benzoquinone to the corresponding p-benzoquinone-substituted bis(NHC)Pd(II) complex 4 (98% yield). The cyclic voltammograms of the Pd(II) complexes 2-4 reveal waves that are attributed to an admix of the anticipated ligand-centred and [Pd(C-NHC)2Br2]-centred processes.  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