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 937 毫秒
1.
A series of binuclear boron compounds supported by Salan(tBu)H4 ligands have been prepared. They are of the general formula Salan(tBu)[B(OR)]2. The compounds are Salean(tBu)(BOR)2 [Salean(tBu) = (N,N′-ethylenebis(3,5-di-tert-butyl-salicylamine)), R = Me (1), SiMe3 (4)], Salban(tBu)(BOR)2[Salban(tBu) = (N,N′-butylenebis(3,5-di-tert-butyl-salicylamine)), R = Me (2), SiMe3 (5)], and Salhan(tBu)(BOR)2 [Salhan(tBu) = (N,N′-hexylenebis(3,5-di-tert-butyl-salicylamine)), R = Me (3)]. All of the compounds were characterized by spectroscopic (1H NMR, 11B NMR, IR) and physical (mp, EA) techniques. Also, 1, 2 and 4 were structurally characterized by single crystal X-ray diffraction studies.  相似文献   

2.
The dioxocyclodiphosph(V)azane cis-[(tBuHN)OP(μ-NtBu)2PO(NHtBu)] reacted with two equivalents of diethylzinc to form the centrosymmetric dimer {[(OPNtBu)2(NtBu)2ZnEt](ZnEt · THF)}2 (1) while under identical conditions, the sulfur and selenium analogues afforded only the monoethylzinc compounds {[(tBuHN)EP(μ-NtBu)2PE(NtBu)](ZnEt · THF)}ES(2), Se (3). To further probe the apparent ligand effects on coordination number and coordination site, cis-[(PhHN)SP(μ-NtBu)2PS(NHtBu)] (5) was synthesized from cis-[ClP(μ-NtBu)2P(NHtBu)] (4) and both were characterized by single-crystal X-ray diffraction. Two equivalents of 5 reacted with diethylzinc to produce the homoleptic, trispirocyclic complex {[(tBuHN)SP(μ-NtBu)2PS(NPh)]2Zn} (6). A second asymmetrically-substituted cyclodiphosph(V)azane, namely [(tBuNH)SP(μ-NtBu)2PNp-tol(NHtBu)] (7), was also synthesized and structurally characterized. In contrast to 5, only one equivalent of this ligand reacted with excess diethylzinc, via its N,N, rather than its N,S side, to afford {[(tBuHN)SP(μ-NtBu)2PNp-tolyl(NtBu)](ZnEt)} (8).  相似文献   

3.
MgMe2 (1) was found to react with 1,4-diazabicyclo[2.2.2]octane (dabco) in tetrahydrofuran (thf) yielding a binuclear complex [{MgMe2(thf)}2(μ-dabco)] (2). Furthermore, from reactions of MgMeBr with diglyme (diethylene glycol dimethyl ether), NEt3, and tmeda (N,N,N′,N′-tetramethylethylenediamine) in etheral solvents compounds MgMeBr(L), (L = diglyme (5); NEt3 (6); tmeda (7)) were obtained as highly air- and moisture-sensitive white powders. From a thf solution of 7 crystals of [MgMeBr(thf)(tmeda)] (8) were obtained. Reactions of MgMeBr with pmdta (N,N,N′,N″,N″-pentamethyldiethylenetriamine) in thf resulted in formation of [MgMeBr(pmdta)] (9) in nearly quantitative yield. On the other hand, the same reaction in diethyl ether gave MgMeBr(pmdta) · MgBr2(pmdta) (10) and [{MgMe2(pmdta)}7{MgMeBr(pmdta)}] (11) in 24% and 2% yield, respectively, as well as [MgMe2(pmdta)] (12) as colorless needle-like crystals in about 26% yield. The synthesized methylmagnesium compounds were characterized by microanalysis and 1H and 13C NMR spectroscopy. The coordination-induced shifts of the 1H and 13C nuclei of the ligands are small; the largest ones were found in the tmeda and pmdta complexes. Single-crystal X-ray diffraction analyses revealed in 2 a tetrahedral environment of the Mg atoms with a bridging dabco ligand and in 8 a trigonal-bipyramidal coordination of the Mg atom. The single-crystal X-ray diffraction analyses of [MgMe2(pmdta)] (12) and [MgBr2(pmdta)] (13) showed them to be monomeric with five-coordinate Mg atoms. The square-pyramidal coordination polyhedra are built up of three N and two C atoms in 12 and three N and two Br atoms in 13. The apical positions are occupied by methyl and bromo ligands, respectively. Temperature-dependent 1H NMR spectroscopic measurements (from 27 to −80 °C) of methylmagnesium bromide complexes MgMeBr(L) (L = thf (4); diglyme (5); NEt3 (6); tmeda (7)) in thf-d8 solutions indicated that the deeper the temperature the more the Schlenk equilibria are shifted to the dimethylmagnesium/dibromomagnesium species. Furthermore, at −80 °C the dimethylmagnesium compounds are predominant in the solutions of Grignard compounds 4-6 whereas in the case of the tmeda complex7 the equilibrium constant was roughly estimated to be 0.25. In contrast, [MgMeBr(pmdta)] (9) in thf-d8 revealed no dismutation into [MgMe2(pmdta)] (12) and [MgBr2(pmdta)] (13) even up to −100 °C. In accordance with this unexpected behavior, 1:1 mixtures of 12 and 13 were found to react in thf at room temperature yielding quantitatively the corresponding Grignard compound 9. Moreover, the structures of [MgMeBr(pmdta)] (9c), [MgMe2(pmdta)] (12c), and [MgBr2(pmdta)] (13c) were calculated on the DFT level of theory. The calculated structures 12c and 13c are in a good agreement with the experimentally observed structures 12 and 13. The equilibrium constant of the Schlenk equilibrium (2 9c ? 12c + 13c) was calculated to be Kgas = 2.0 × 10−3 (298 K) in the gas phase. Considering the solvent effects of both thf and diethyl ether using a polarized continuum model (PCM) the corresponding equilibrium constants were calculated to be Kthf = 1.2 × 10−3 and Kether = 3.2 × 10−3 (298 K), respectively.  相似文献   

4.
Cadmium(II) complexes of 3-hydroxypicolinic acid, namely [CdI(3-OHpic)(3-OHpicH)(H2O)]2 (1), [Cd(3-OHpic)2(H2O)2] (2) and [Cd(3-OHpic)2]n (3) were prepared and characterized by spectroscopic methods (IR, NMR) and their molecular and crystal structures were determined by X-ray crystal structure analysis. Complexes 1 and 2 were prepared in similar reaction conditions using different cadmium(II) salts: cadmium(II) iodide and cadmium(II) acetate dihydrate, respectively, while 3 was prepared by recrystallization of 2 from N,N-dimethylformamide solution. Various coordination modes of 3-OHpicH in 13 were established in the solid state: bidentate N,O-chelated mode in 1 and 2, monodentate mode through the carboxylate O atom from zwitterionic ligand in 1 and bidentate N,O-chelated and bridging mode in 3. In the DMF solution of all prepared complexes, only monodentate mode of 3-OHpicH binding to cadmium(II) through the carboxylate O atom was established by 1H, 13C, 15N and 113Cd NMR spectroscopy.  相似文献   

5.
Reactions of Ru3(CO)12 with diphosphazane monoselenides Ph2PN(R)P(Se)Ph2 [R = (S)-∗CHMePh (L4), R = CHMe2 (L5)] yield mainly the selenium bicapped tetraruthenium clusters [Ru44-Se)2(μ-CO)(CO)8{μ-P,P-Ph2PN(R)PPh2}] (1, 3). The selenium monocapped triruthenium cluster [Ru33-Se)(μsb-CO)(CO)72-P,P-Ph2PN((S)-∗CHMePh)PPh2}] (2) is obtained only in the case of L4. An analogous reaction of the diphosphazane monosulfide (PhO)2PN(Me)P(S)(OPh)2 (L6) that bears a strong π-acceptor phosphorus shows a different reactivity pattern to yield the triruthenium clusters, [Ru33-S)(μ3-CO)(CO)7{μ-P,P-(PhO)2PN(Me)P(OPh)2}] (9) (single sulfur transfer product) and [Ru33-S)2(CO)52-P,P-(PhO)2PN(Me)P(OPh)2}{μ-P,P-(PhO)2PN(Me)P(OPh)2}] (10) (double sulfur transfer product). The reactions of diphosphazane dichalcogenides with Ru3(CO)12 yield the chalcogen bicapped tetraruthenium clusters [Ru44-E)2(μ-CO)(CO)8{μ-P,P-Ph2PN(R)PPh2}] [R = (S)-∗CHMePh, E = S (6); R = CHMe2, E = S (7); R = CHMe2, E = Se (3)]. Such a tetraruthenium cluster [Ru44-S)2(μ- CO)(CO)8{μ-P,P-(PhO)2PN(Me)P(OPh)2}] (11) is also obtained in small quantities during crystallization of cluster 9. The dynamic behavior of cluster 10 in solution is probed by NMR studies. The structural data for clusters 7, 9, 10 and 11 are compared and discussed.  相似文献   

6.
The reaction of N9,N9′-(tri or tetramethylene)-bisadenines (Ade2Cx; x = 3 or 4) in HCl 2 M at 50 °C with MCl2 · 2H2O [M = Zn(II), Cd(II)] yields outer sphere compounds like the previously described [(H-Ade)2C3][ZnCl4] · H2O (3) and [(H-Ade)2C3]2[Cd2Cl8(H2O)2] · 4H2O (4) for Ade2C3 and the new {[(H-Ade)2C4][Cd2Cl6(H2O)2] · 2H2O}n (5) for Ade2C4. On the other hand, only in case of Zn(II) complexes by changing [HCl] to 0.1 M, the inner sphere compounds [H-(Ade)2C3(ZnCl3)] (6) and [H-(Ade)2C4(ZnCl3)] · 1.5H2O (7) are obtained. X-ray diffraction study of compound 6, which represents the first inner sphere complex with a N9,N9′-bisadenine, shows a zwitterionic form with one adenine ring protonated at N(1) while the other ring is coordinated via N(7) to a ZnCl3 moiety as in other alkyl-adenine derivatives. In addition, with Ade2C4, is also possible to obtain another inner sphere complex: [(H-Ade)2C4(ZnCl3)2] · 3H2O (8).  相似文献   

7.
Mononuclear complexes of the type, M(CO)4[Se2P(OR)2] (M = Mn, R = iPr, 1a; Et, 1b; M = Re, R = iPr, 3a; Et, 3b) can be prepared from either [-Se(Se)P(OiPr)2]2 (A) or [Se{-Se(Se)P(OEt)2}2] (B) with M(CO)5Br. O,O′-dialkyl diselenophosphate ([(RO)2PSe2]-, abbreviated as dsep) ligands generated from A and B act as a chelating ligand in these complexes. Upon refluxing in acetonitrile, these mononuclear complexes yield dinuclear complexes with a general formula of [M2(CO)6{Se2P(OR)2}2] (M = Mn, R = iPr, 2a; Et, 2b; M = Re, R = iPr, 4a; Et, 4b). Dsep ligands display a triconnective, bimetallic bonding mode in the dinuclear compounds and this kind of connective pattern has never been identified in any phosphor-1,1-diselenoato metal complexes. Compounds 2b, 3b, and 4 are structurally characterized. Compounds 2b and 3b display weak, secondary Se?Se interactions in their lattices.  相似文献   

8.
The reaction of bis(2-pyridylmethyl)amine (II) with t-butylamine and dimethylzinc gives the heteroleptic [(MeZn)2{μ-N(H)tBu}{μ-N(CH2Py)2}] (1). Stoichiometric alcoholysis of 1 with methanol leads to the exchange of the μ-N(H)tBu moiety. Almost quantitatively the corresponding methoxide [(MeZn)2(μ-OMe){μ-N(CH2Py)2}] (2) is formed. Alternatively bis(alkylzinc)methoxide-bis(2-pyridylmethyl)amides (Alkyl = methyl (2), bis(trimethylsilyl)methyl) (3)) are also accessible by direct zincation of bis(2-pyridylmethyl)amine (II) and methanol with dialkylzinc regardless of the bulkiness of the alkyl groups. Extensive DFT calculations on the alcoholysis mechanism reveal the preferential insertion of methanol into a zinc amide bond rather than the cleavage of zinc carbon bonds. An intermediate with a Zn[μ-(MeO?H?NHR)]Zn functionality is predicted. Aminolyis of 1 with t-butylamine leads to intermediates with Zn[μ-(RNH ? H ? NHR)]Zn functionalities, respectively. We were able to detect the latter by 1H NMR spectroscopy. The aminolysis of 1 with an excess of phenylamine results in a partial decomposition of the complex leading to the hexanuclear amide [{Zn(μ-N(H)Ph)}{MeZn(μ-N(H)Ph)}2{μ-N(CH2Py)2}]2 (4). Compound 2 is able to cleave silicon grease when dissolved in t-butylamine yielding [(MeZn)2{μ-N(CH2Py)2}2Zn{μ-(OMe2Si)2O}] (5). The X-ray structures of complexes 1-5 are discussed.  相似文献   

9.
The reaction of Fe2(CO)9 with Bi(OSiMe2tBu)3 gave soluble [(CO)4FeBi(OSiMe2tBu)]2 (1) in moderate yield whereas in case of Bi(OtBu)3 used as starting material both [(CO)4FeBi(OtBu)]n (2) and the bismuth-iron cluster [(CO)3FeBi3(OtBu)4{OCO(OtBu)}]2 (3) were isolated. The latter forms upon insertion of CO2, released during reaction of diiron nonacarbonyl with bismuth tert-butoxide, into a Bi-OtBu bond. The compounds were characterized by IR and 1H NMR spectroscopy as well as thermogravimetric analyses. Additionally, the molecular structures of compounds 1 and 3 were elucidated by single crystal X-ray diffraction. The core structure of [(CO)4FeBi(OSiMe2tBu)]2 (1) is build up by a four-membered Bi2Fe2 ring whereas [(CO)3FeBi3(OtBu)4{OCO(OtBu)}]2 (3) is composed of two tetrahedral FeBi3 cluster cores that dimerise via bridging -OCO(OtBu) ligands. Analysis of the TGA residues by PXRD revealed that compound 2 is the best precursor for multiferroic BiFeO3 among the compounds studied here, although Bi25FeO39 was detected as minor impurity.  相似文献   

10.
The reaction of the cluster salt K4[Re4Te4(CN)12]·5H2O with NdCl3·6H2O was studied in either an acidic medium (HCl) or in a water solution in the presence of the following organic agents: hexafluoroacetylacetonate, 2,2′-bipyridine or 1,10-phenanthroline (phen). The crystal structures of four new compounds have been solved by single crystal X-ray diffraction analysis: (H)[{Nd(H2O)5}{Re4Te4(CN)12}]·5.5H2O (1) (space group P21/c, framework structure), K2[{Nd(H2O)7}2{Re4Te4(CN)12}2]·8H2O (2) (space group С2/c, isolated structure), K0.5H0.5[{Nd(H2O)5}{Re4Te4(CN)12}]·3H2O (3) (space group Сmcm, layered structure) and (phenH)[{Nd(H2O)2(phen)2}{Re4Te4(CN)12}]·11H2O (4) (space group С2/c, chain structure). 1,10-Phenanthroline was found to have been incorporated into the structure of compound 4, whilst hexafluoroacetylacetonate and 2,2′-bipyridine did not enter the structures of 2 and 3. It was shown that the structures of compounds 2-4 differ dramatically from that found for compound 1, which was obtained in the absence of the organic agents.  相似文献   

11.
The use of succinamic acid (H2sucm) in CuII/N,N′,N″-donor [2,2′:6′,2″-terpyridine (terpy), 2,6-bis(3,5-dimethylpyrazol-1-yl)pyridine (dmbppy)] reaction mixtures yielded compounds [Cu(Hsucm)(terpy)]n(ClO4)n (1), [Cu(Hsucm)(terpy)(MeOH)](ClO4) (2), [Cu2(Hsucm)2(terpy)2](ClO4)2 (3), [Cu(ClO4)2(terpy)(MeOH)] (4), [Cu(Hsucm)(dmbppy)]n(NO3)n·3nH2O (5.3nH2O), and [CuCl2(dmbppy)]·H2O (6·H2O). The succinamate(−1) ligand exists in four different coordination modes in the structures of 13 and 5, i.e., the μ2OO′:κO″ in 1 and 5 which involves asymmetric chelating coordination of the carboxylato group and ligation of the amide O-atom leading to 1D coordination polymers, the μ22OO′ in 3 which involves asymmetric chelating and bridging coordination of the carboxylato group, and the asymmetric chelating mode in 2. The primary amide group, either coordinated in 1 and 5, or uncoordinated in 2 and 3, participate in hydrogen bonding interactions, leading to interesting crystal structures. Characteristic IR bands of the complexes are discussed in terms of the known structures and the coordination modes of the Hsucm ligands. The thermal decomposition of complex 5·3nH2O was monitored by TG/DTG and DTA measurements.  相似文献   

12.
N,N,N′,N′-Tetramethylmethanediamine (1a), N,N,N′,N′-tetramethylethanediamine (1b), N,N,N′,N′-tetramethyl-1,3-propanediamine (1c), and N,N,N′,N′-tetramethyl-1,6-hexanediamine (1d) were reacted at 25 °C with 1,1,1,5,5,5-hexafluoro-2,4-pentanedione (2a), 2,2-dimethyl-6,6,7,7,8,8,8-heptafluoro-3,5-octanedione (2b), 2-thenoyltrifluoroacetone (2c), and 4,4,4-trifluoro-1-(2-furyl)-1,3-butanedione (2d) to form the ionic adducts 3-18. 1,4,7,10-Tetraazacyclododecane (1e) reacted at 25 °C with β-diketones (2a-d) and 1,1,1-trifluoro-2,4-pentanedione (2e) to give ionic solids 19-23 in good yields. Some of the products are liquid at 25 °C and are thermally stable over long liquid ranges as determined by thermal gravimetric analyses. Single-crystal X-ray structure determinations show that compounds 9 and 21 crystallize in the monoclinic space groups P2(1)/c and P2(1)/n, respectively. All the new compounds were characterized by 1H, 19F and 13C NMR, electrospray MS and/or elemental analyses.  相似文献   

13.
Schiff base N,N′-bis(salicylidene)-p-phenylenediamine (LH2) complexed with Pt(en)Cl2 and Pd(en)Cl2 provided [Pt(en)L]2 · 4PF6 (1) and Pd(Salen) (2) (Salen = N,N′-bis(salicylidene)-ethylenediamine), respectively, which were characterized by their elemental analysis, spectroscopic data and X-ray data. A solid complex obtained by the reaction of hexafluorobenzene (hfb) with the representative complex 1 has been isolated and characterized as 3 (1 · hfb) using UV–Vis, NMR (1H, 13C and 19F) data. A solid complex of hfb with a reported Zn-cyclophane 4 has also been prepared and characterized 5 (4 · hfb) for comparison with complex 3. The association of hfb with 1 and 4 has also been monitored using UV–Vis and luminescence data.  相似文献   

14.
Ligand effects on the catalytic activity [and norbornene (NBE) incorporation] for both ethylene polymerization and ethylene/NBE copolymerization using half-titanocenes (titanium half-sandwich complexes) containing ketimide ligand of type Cp′TiCl2[NC(R1)R2] [Cp′ = Cp (1), C5Me5 (Cp, 2); R1,R2 = tBu,tBu (a), tBu,Ph (b), Ph,Ph (c)]-methylaluminoxane (MAO) catalyst systems have been investigated. CpTiCl2[NC(tBu)Ph] (1b) CpTiCl2(NCPh2) (1c), and CpTiCl2(NCPh2) (2c) were prepared and identified; the structure of CpTiCl2(NCPh2) (2c) was determined by X-ray crystallography. The catalytic activity for ethylene polymerization increased in the order: 1a > 1b > 1c, suggesting that an electronic nature of the ketimide ligand affects the activity. However, molecular weight distributions for resultant (co)polymers prepared by 1b,c and by 2c-MAO catalyst systems were bi- or multi-modal, suggesting that the ketimide substituent plays a key role in order for these (co)polymerizations to proceed with single catalytically-active species. CpTiCl2(NCtBu2) (1a) exhibited both remarkable catalytic activity and efficient NBE incorporation for ethylene/NBE copolymerization.  相似文献   

15.
The syntheses and crystal structures of four new uranyl complexes with [O,N,O,N′]-type ligands are described. The reaction between uranyl nitrate hexahydrate and the phenolic ligand [(N,N-bis(2-hydroxy-3,5-dimethylbenzyl)-N′,N′-dimethylethylenediamine)], H2L1 in a 1:2 molar ratio (M to L), yields a uranyl complex with the formula [UO2(HL1)(NO3)] · CH3CN (1). In the presence of a base (triethylamine, one mole per ligand mole) with the same molar ratio, the uranyl complex [UO2(HL1)2] (2) is formed. The reaction between uranyl nitrate hexahydrate and the ligand [(N,N-bis(2-hydroxy-3,5-di-t-butylbenzyl)-N′,N′-dimethylethylenediamine)], H2L2, yields a uranyl complex with the formula [UO2(HL2)(NO3)] · 2CH3CN (3) and the ligand [N-(2-pyridylmethyl)-N,N-bis(2-hydroxy-3,5-dimethylbenzyl)amine], H2L3, in the presence of a base yields a uranyl complex with the formula [UO2(HL3)2] · 2CH3CN (4). The molecular structures of 14 were verified by X-ray crystallography. The complexes 14 are zwitter ions with a neutral net charge. Compounds 1 and 3 are rare neutral mononuclear [UO2(HLn)(NO3)] complexes with the nitrate bonded in η2-fashion to the uranyl ion. Furthermore, the ability of the ligands H2L1–H2L4 to extract the uranyl ion from water to dichloromethane, and the selectivity of extraction with ligands H2L1, H3L5 (N,N-bis(2-hydroxy-3,5-dimethylbenzyl)-3-amino-1-propanol), H2L6 · HCl (N,N-bis(2-hydroxy-5-tert-butyl-3-methylbenzyl)-1-aminobutane · HCl) and H3L7 · HCl (N,N-bis(2-hydroxy-5-tert-butyl-3-methylbenzyl)-6-amino-1-hexanol · HCl) under varied chemical conditions were studied. As a result, the most efficient and selective ligand for uranyl ion extraction proved to be H3L7 · HCl.  相似文献   

16.
The complexes [Rh(CO)(PPh3){Ph2PNP(O)Ph2-P,O}] (3), [Rh(CO)2{Ph2P(Se)NP(Se)Ph2-Se,Se′}] (5), and [Rh(CO)(PPh3){Ph2P(Se)NP(Se)Ph2-Se,Se′}] (6), were synthesised by stepwise reactions of CO and PPh3 with [Rh(cod){Ph2PNP(O)Ph2-P,O}] (2) and [Rh(cod){Ph2P(Se)NP(Se)Ph2-Se,Se′}] (4), respectively. The complexes 3, 5 and 6 have been studied by IR, as well as 1H and 31P NMR spectroscopy. The ν(CO) bands of complexes 3 and 6 appear at approximately 1960 cm−1, indicating high electron density at the RhI centre. The structure of complexes 3 and 6 has been determined by X-ray crystallography, and the 31P NMR chemical shifts have been resolved via low temperature NMR experiments. Both complexes exhibit square planar geometry around the metal centre, with the five-membered ring of complex 3 being almost planar, and the six-membered ring of complex 6 adopting a slightly distorted boat conformation. The C-O bond of the carbonyl ligand is relatively weak in both complexes, due to strong π-back donation from the electron rich RhI centre. The catalytic activity of the complexes 2, 3 and 6 in the hydroformylation of styrene has been investigated. Complexes 2 and 3 showed satisfactory catalytic properties, whereas complex 6 had effectively no catalytic activity.  相似文献   

17.
Alkyl aluminum N,N′-dimethyloxalamidates R4Al2(dmoa) (1, R = Me; 2, R = Et; 3, R = iBu; 4, R = tBu) (dmoa-H2 = N,N′-dimethyloxalamide) have been prepared and characterized. Molecular structures of the compounds 1 and 4 have been determined by X-ray crystallography. The centrosymmetric molecules of the compounds consist of one N,N′-dimethyloxalamidate unit bonded to two four-coordinated aluminum atoms. Each of the aluminum atoms is bonded to two alkyl groups, and oxygen and nitrogen atoms originating from two different amidate groups. A skeleton framework of the molecules of 1 and 4 consists of two fused AlNOC2 heterocyclic rings, which are flat and positioned in one plane. It was shown that compounds 1-3 were initiators in a process of ring opening polymerization (ROP) of ε-caprolactone. The compound 4 exhibited low activity in ROP.  相似文献   

18.
The rhenium(I) carbonyl halide (X = Cl and Br) complexes, [ReX(CO)3{H2(py)L2}] (1a, 1b) and [ReX(CO)3{H2(Fc)L2}] (2a, 2b), of the ligands derived from 2-acetylpyridine and ferrocenyl carbaldehyde derivatives of 2-hydroxybenzoic acid hydrazide [H2(py)L2 and H2(Fc)L2, respectively] have been prepared in good yield. The complexes have been characterized by elemental analysis, MS, IR, UV-Vis and 1H NMR spectroscopic methods and their structures have been elucidated by X-ray diffraction. The ligand forms a five-membered chelate ring but in H2(py)L2 it is Npyridine,N′-bidentate while it is O,N-bidentate in H2(Fc)L2 complexes.Reaction of complex 1a with copper(II) nitrate yields the unexpected aqua complex [Re{H(py)L2}(H2O)(CO)3] (3) where the ligand is monodeprotonated but maintains the coordination mode observed in 1a, as shown by X-ray diffraction. However, reaction of 1b with glycine yields a conformational polymorph of the original compound, 1b′. The X-ray study shows that the orientation of the O-H phenol group against the carbonyl amide group is the main difference.  相似文献   

19.
Five transition metal compounds containing arenesulfonates and 4,4′-bipy ligands, namely [Zn2(N,N′-4,4′-bipy)(N-4,4′-bipy)2(H2O)8](bpds)2 · 5H2O (1), [Ag2(N,N′-4,4′-bipy)2(bpds)] (2), [Cd(N,N′-4,4′-bipy)(H2O)4]2(4-abs)4 · 5H2O (3), [Cu(N,N′-4,4′-bipy) (O-bs)2(H2O)2] · 4H2O (4), and [Zn(N,N′-4,4′-bipy)2(H2O)2](4,4′-bipy)(bs)2 · 4H2O (5) (4,4′-bipy = 4,4′-bipyridine, bpds = 4,4′-biphenyldisulfonate, 4-abs = 4-aminobenzenesulfonate, bs = benzenesulfonate), have been synthesized and characterized by X-ray single crystal diffraction, elemental analyses and TG analyses, in order to investigate the coordination chemistry of arenesulfonates and 4,4-bipy, as well as to construct novel coordination frameworks via mixed-ligand strategy. Compounds 2, 4 and 5 could be obtained via hydrothermal or aqueous reactions. Compound 1 forms a binuclear octahedral metal complex. Compounds 24 form polymeric chains. Compound 5 consists of 2D square grids with one intercalated 4,4′-bipy molecule. Weak Ag–Ag interactions are observed in compound 2. These complexes show great structural varieties and there are three different coordination modes observed for both the 4,4′-bipy and the sulfonate ligands.  相似文献   

20.
A new tetradentate imidazolate ligand 1,1′,1″,1′′′-(2,2′,4,4′,6,6′-hexamethylbiphenyl-3,3′,5,5′-tetrayl)tetrakis(methylene)(1H-imidazole) (L) and four Ag(I)/Cu(I) coordination polymers, namely [(MCN)3L]n (1: M=Ag; 2: M=Cu), and [(MSCN)2L]n (3: M=Ag; 4: M=Cu) are described. All four new coordination polymers were fully characterized by infrared spectroscopy, elemental analysis and single-crystal X-ray diffraction. Compound 1 features a 3D supramolecular framework constructed by 1D chains through inter-chain Ag-N(CN) and inter-layer Ag-N(L) weak interactions with an uninodal 66 topology. Complex 2 presents a 3D framework characterized by a tetranodal (3,4)-connected (3·4·5·102·11)(3·4·5·6·7·9)(3·6·7)(6·102) topology. Complexes 3 and 4 are isostructural, and both have a 3D network of trinodal 4-connected (4·85)2(42·82·102)(42·84)2 topology. The luminescent properties for these compounds in the solid state as well as the possible ferroelectric behavior of 1 are discussed.  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