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 93 毫秒
1.
Reported are multi‐component one‐pot syntheses of chiral complexes [M(LROR′)Cl2] or [M(LRSR′)Cl2] from the mixture of an N‐substituted ethylenediamine, pyridine‐2‐carboxaldehyde, a primary alcohol or thiol and MCl2 utilizing in‐situ formed cyclized Schiff bases where a C?O bond, two stereocenters, and three C?N bonds are formed (M=Zn, Cu, Ni, Cd; R=Et, Ph; R′=Me, Et, nPr, nBu). Tridentate ligands LROR′ and LRSR′ comprise two chiral centers and a hemiaminal ether or hemiaminal thioether moiety on the dipicolylamine skeleton. Syn‐[Zn(LPhOMe)Cl2] precipitates out readily from the reaction mixture as a major product whereas anti‐[Zn(LPhOMe)Cl2] stays in solution as minor product. Both syn‐[Zn(LPhOMe)Cl2] and anti‐[Zn(LPhOMe)Cl2] were characterized using NMR spectroscopy and mass spectrometry. Solid‐state structures revealed that syn‐[Zn(LPhOMe)Cl2] adopted a square pyramidal geometry while anti‐[Zn(LPhOMe)Cl2] possesses a trigonal bipyramidal geometry around the Zn centers. The scope of this method was shown to be wide by varying the components of the dynamic coordination assembly, and the structures of the complexes isolated were confirmed by NMR spectroscopy, mass spectrometry, and X‐ray crystallography. Syn complexes were isolated as major products with ZnII and CuII, and anti complexes were found to be major products with NiII and CdII. Hemiaminals and hemiaminal ethers are known to be unstable and are seldom observed as part of cyclic organic compounds or as coordinated ligands assembled around metals. It is now shown, with the support of experimental results, that linear hemiaminal ethers or thioethers can be assembled without the assistance of Lewis acidic metals in the multi‐component assembly, and a possible pathway of the formation of hemiaminal ethers has been proposed.  相似文献   

2.
Reaction of [MnBr(CO)3L] [L = Ph2POCH2CH2OPPh2, L1 , {(CH3)2CH}2POCH2CH2OP{CH(CH3)2}2, L2 ] with AgO3SCF3 and AgO2CCF3 in dichloromethane afforded the new complexes [Mn(O3SCF3)(CO)3L] and [Mn(O2CCF3)(CO)3L], respectively. Substitution of O3SCF3 resulted in the new species [Mn(SCN)(CO)3L], [Mn(NCCH3)(CO)3L](O3SCF3) and, in the case of L2 , [Mn(CN)(CO)3L2]. By contrast, any attempt to displace the O2CCF3 ligand in the same way was unsuccessful. After maintaining for some days the complex [Mn(CH3CN)(CO)3L1](O3SCF3) in dichloromethane at room temperature, the new complex [MnCl(CO)3L1] was formed. All the new complexes were characterized by elemental analysis, mass spectrometry and IR and NMR spectroscopies. In the case of [Mn(O3SCF3)(CO)3L1], [Mn(O2CCF3) (CO)3L1], [MnCl(CO)3L1], [Mn(CH3CN) (CO)3L2] (O3SCF3), [Mn(CN)(CO)3L2] and [Mn(O2CCF3)(CO)3L2], together with the previously synthesized complex [MnBr(CO)3L2], suitable crystals for X‐ray structural analysis were isolated. In all of them the Mn atom adopts six‐coordination by bonding to the three CO ligands, the two P atoms of L and either one C atom (CN), one oxygen atom (O2CCF3, O3SCF3), one N atom (CH3CN, SCN) or the halogen atom (Cl, Br).  相似文献   

3.
Electron-rich, half-sandwich complexes of the type CnRnML2 or CnRnMLL′ are built up of an aromatic five- or six-membered ring, a d8-metal, and either a pair of two-electron donors or an equivalent chelating ligand. Such complexes behave like Lewis bases and react with a wide variety of electrophiles, El or ElX, to form products with a new metal-element bond. According to their reactivity they are comparable to the Vaska-type compounds. Certain of the products obtained after addition of the electrophile undergo interesting subsequent reactions in which, for example, metal complexes containing molecules that are unstable in the free state, such as CS, CSe, CH2S, CH2Se, CH2Te, CH3CHS, CH3CHSe, CH2?C?S, CH2?C?Se, and CH2?C?Te are formed. Moreover, cycloadditions as well as reactions with coordinatively unsaturated transition-metal compounds which result in formation of heterometal binuclear complexes demonstrate that the metal bases CnRnML2 and CnRnMLL′ are valuable synthetic building blocks. Furthermore, very recent investigations have indicated links between metal basicity and the problem of C? H activation.  相似文献   

4.
The carboxylate compounds [Ti(η5‐C5H5)(η5‐C5H4{CMe2(CH2CH2CH?CH2)})(O2CCH2SXyl)2] (2; Xyl = 3,5‐Me2C6H3) and [Ti(η5‐C5H5)(η5‐C5H4{CMe2(CH2CH2CH?CH2)})(O2CCH2SMesl)2] (3; Mes 1 = 2,4,6‐Me3C6H2) were synthesized by the reaction of [Ti(η5‐C5H5)(η5‐C5H4{CMe2(CH2CH2CH?CH2)})Cl2] (1) with 2 equivalents of xylylthioacetic acid or mesitylthioacetic acid, respectively. Compounds 2 and 3 were characterized by spectroscopic methods. The cytotoxic activity of 1–3 was tested against human tumor cell lines from four different histogenic origins—8505C (anaplastic thyroid cancer), DLD‐1 (colon cancer) and the cisplatin sensitive A253 (head and neck cancer) and A549 (lung carcinoma)—and compared with those of the reference complex [Ti(η5‐C5H5)2Cl2] (R1) and cisplatin. Surprisingly, the cytotoxic activities of the carboxylate derivatives were lower than those of their corresponding dichloride analogue (1). However, complexes 1–3 were more active than titanocene dichloride against all the studied cells with the exception of complex 2 against A253 and A549 cell lines. DNA‐interaction tests were also carried out. Solutions of all the studied complexes were treated with different concentrations of fish sperm DNA, observing modifications of the UV spectra with intrinsic binding constants of 2.99 × 105, 2.45 × 105, and 2.35 × 105 M ?1 for 1–3. Structural studies based on density functional theory calculations of 2 and 3 were also carried out. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

5.
Cyclooctyne is the smallest unsubstituted cycloalkyne that can be isolated in the free state and it is more reactive than acyclic alkynes towards transition metal complexes. Smaller cycloalkynes such as cycloheptyne, cyclohexyne, benzyne and cyclopentyne, which are transient molecules in the free state, can be stabilized by coordination either to mononuclear, electron-rich, transition metal-containing fragments, e.g. [ZrCp2(PMe3)] and M(PR3)2 (M = Ni, Pt), or by formation of dinuclear or polynuclear metal complexes, e.g. [Os3H2(CO)9(C6H4)] and [Pt2(μ-PPh2)(μ-C5H6)(PPh3)3]. The alkynes can donate between two and four π-electrons to the metal centers, the higher number being favored for the early transition metals. The metal-cycloalkyne and metal-aryne bonds in the mononuclear complexes readily insert molecules containing C?O, C?C, C?C and C?N bonds, a feature that may be useful in organic synthesis. The highly unsaturated species 1, 4-benzdiyne acts as a bridging ligand between two metal centers in [{Ni(Cy2PCH2CH2PCy2)}2(μ-C6H2)].  相似文献   

6.
To examine the interaction of uranyl with nitrogen containing groups of humic substances, the model complexes [UO2(H2O)4LN]2+, LN = NH2CH3, N(CH3)3, and NC5H5 in aqueous solution were studied computationally with an all‐electron relativistic density functional method. Results are compared with the corresponding penta‐aqua complex of uranyl. Although pyridine coordinates with about the same strength as L = H2O, methylamine binds ~10 kJ mol?1 stronger and trimethylamine ~40 kJ mol?1 weaker than a fifth aqua ligand. Yet, each of these ligands LN donates about the same amount of charge to uranyl as L = H2O. U? N bonds are ~10 pm longer than the U? O bonds of the aqua ligands. From the present model results, one does not expect that, when compared with carboxyl groups, monodentate N‐containing functional groups contribute significantly to uranyl complexation by humic substances. © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

7.
Abstract

The ligand chemistry of telluroethers, halotellurium ligands, and polytellurides has received good attention in the last decade. Tellurium-containing species have been used to design clusters. In the recent past the ligation of di and tri-telluroethers (including bis(4-methoxyphenyltelluro)methane) has been studied. Hybrid organotellurium ligands, N-[2-(4-methoxyphenyltelluro)propyl]phthalimid (L 1 ), 2-(4-ethoxyphenyltelluromethyl)-tetrahydro-2H-pyran (L 2 ), 2-(2-{4-ethoxyphenyl} telluroethyl)-1,3-dioxane (L 3 ), N-{2-(4-methoxyphenyltelluro)ethyl}morpholine (L 4 ), N-{2-(4-methoxyphenyltelluro)ethyl}-pyrrolidine (L 5 ), bis{2-(pyrrolidine-N-yl)ethyl}telluride (L 6 ), 1-(4-methoxyphenyltelluro)-2-[3-(6-methyl-2-pyridyl) propoxy]ethane (L 7 ), and 2-[2-(4-methoxyphenyltelluro)ethyl]thiophene (L 8 ) have been designed recently and studied for their complexation reactions. The (Te, N) and (N, Te, N) ligands, L 5 and L 6 , coordinate with Hg(II) through Te and N both, but the bonding with N is some what weak. The morpholine nitrogen of L 4 does not coordinate with Pd(II) or Pt(II) along with Te. The L 7 behaving as a (Te, N) ligand has formed 20-membered metallomacrocycle ring with Pt(II). Tellurated Schiff bases 4-MeOC6H4TeCH2CH2N═C(CH3)C6H4-2-OH (L 9 ) and 2-HO-C6H4-(CH3)C═NCH2CH2TeCH2CH2N═C(CH3)C6H4-2-OH (L 10 ) and their reduction products 4-MeOC6H4TeCH2CH2NHCH(CH3)C6H4-2-OH (L 11 ) and 2-HO-C6H4-(CH3)CHNHCH2CH2TeCH2CH2NHCH(CH3)C6H4-2-OH (L 12 ) respectively have been synthesized and studied for ligation behaviour. The L 9 on reaction with the [Ru(p-cymene)Cl2]2 results in [Ru(p-cymene)(4-MeOC6H4TeCH2CH2NH2)Cl]Cl · H2O whereas in the reaction of L 10 with [Ru(p-cymene) Cl2]2, p-cymene ligand is lost resulting in [RuCl(L 10 -H)]. The recent developments, particularly designing of L 1 to L 12 and their ligand chemistry, are reviewed in the present paper.  相似文献   

8.
Abstract

The synthesis of Ti(iso)Cl2 (iso = the dianion of 2, 2′-ethylidenebis(4, 6-di-rert-butylphenol)) is described. Metathesis reactions of this complex with Grignard reagents, as well as alkali metal salts, yielded Ti(iso)X2 (X = CH3, CH, Ph, CH2SiMe3, OCMe3, or NMe2). Reactions of the Ti-C bond in Ti(iso)(CH3)2 toward halogens, active hydrogen compounds, and acetone were studied. In addition. Ti(iso)(X)(Y) (X = CI or CH3; Y = OC6H2-2, 6-tBu2-4-Me) could be prepared with the formation of only one coordination isomer. The new complexes have been thoroughly characterized by 1H and 13C NMR spectroscopies. The solid state structure of Ti(iso)Cl2 was determined via single crystal X-ray diffraction methods. The complex is monomeric, with approximately tetrahedral geometry about the titanium ion. The structure of Ti(iso)Cl2 is compared to that of Ti(ultra)Cl2 (ultra = the dianion of 2, 2′-mefhylenebis(6-tert-butyl-4-mefhylphenol)), which was redetermined to greater precision.  相似文献   

9.
The preparation of tetraethylammonium slats with following anions is described: [TiBrCl5]2?,[TiBr5Cl]2?,[TiCl4Br · CH3CN]?, [TiBr4Cl · CH3CN]?, [TiBr5 · CH3CN]? und TiF. The reaction mechanisms is discussed. TiF forms fluorine-bridges giving a polymeric anion. Chlorofluorotitanates(IV) could not be prepared. Mixed halide complexes of Ti(IV), Nb(V) and Ta(V) are compared with analogous complexes of Sn(IV), Pb(IV) and Sb(V).  相似文献   

10.
Bibracchial lariat ethers L3 and L4, derived from the condensation of N,N′-bis(2-aminobenzyl)-1,10-diaza-15-crown-5 or N,N′-bis(2-aminobenzyl)-4,13-diaza-18-crown-6 with salicylaldehyde, form binuclear complexes with Co(II), Ni(II), Cu(II) and Zn(II). Our studies show that the different denticity and crown moiety size of the two related receptors give rise to important differences on the structures of the corresponding complexes. Single crystal X-ray diffraction analysis shows that the [Ni2(L3)(H2O)2]2+ and [Cu2(L3)(NO3)]+ complexes constitute a rare example in which an oxygen atom of the crown moiety is bridging the two six coordinate metal ions. In contrast, none of the oxygen atoms of the crown moiety is acting as a bridging donor atom in the [Co2(L4)(CH3CN)2]2+, [Cu2(L4)]2+ and [Zn2(L4)]2+ complexes. This is attributed to the larger size the crown moiety and the higher denticity of L4 compared to L3. In [Co2(L4)(CH3CN)2]2+ the metal ions show a distorted octahedral coordination, while in the Cu(II) and Zn(II) analogues the metal ions are five-coordinated in a distorted trigonal bipyramidal environment. In [Cu2(L3)(NO3)]+ the coordinated nitrate anion acts as a bidentate bridging ligand, which results in the formation of a 1D coordination polymer.  相似文献   

11.
Pyromellitic diimide and naphthalene diimide moieties were incorporated into hemilabile phosphanylalkyl thioether ligands. These ligands reacted with [Cu(CH3CN)4]PF6 and [Rh(NBD)Cl]2 (NBD=norbornadiene) by the weak‐link approach to form condensed intermediates. Upon reaction of each diimide ligand with these transition‐metal precursors, the two diimide units became cofacially aligned within a supramolecular macrocyclic architecture. The introduction of ancillary ligands to each of these condensed intermediates caused the weak thioether–metal bonds to break, thus generating a large macrocycle in which the distance between diimide units is significantly larger than for the condensed intermediates. The two RhI cationic condensed intermediates were characterized by single‐crystal X‐ray diffraction studies, and the electrochemical activity of these macrocycles was demonstrated with the naphthalene diimide–CuI macrocycles.  相似文献   

12.
Two heterometallic trinuclear complexes of macrocyclic oxamide [Co(Ni L1 )2 L2 (H2O)] · 3H2O ( 1 ) and [Mn(Ni L1 )2 L2 (H2O)] · 0.5CH3OH · 1.5H2O ( 2 ) (H2 L1 = 2,3‐dioxo‐5,6,14,15‐dibenzo‐1,4,8,12‐tetraazacyclopentadeca‐7,13‐diene, H2 L2 = 5‐sulfosalicylic acid) were synthesized and structurally characterized by elemental analysis, IR spectroscopy, and X‐ray diffraction. Single‐crystal X‐ray analyses reveal that both the complexes contain discrete neutral trinuclear [(Ni L1 )2M L2 (H2O)] (for 1 and 2 , M = Co, Mn, respectively) moieties. The structures of 1 and 2 have oxamido‐bridged trinuclear [MIINiII2] units and consist of one‐dimensional chains formed by strong intermolecular hydrogen bonds. Furthermore, the magnetic properties of complex 1 were investigated and discussed in detail.  相似文献   

13.
The recognition of the fundamental contributions by G. A. Olah on the elucidation of the structure of nonclassical carbocations, in the form of the award of the Nobel prize for chemistry, has recently emphasized the importance of electron-deficient bonds in the understanding of chemical bonding in organic chemistry. In the field of coordination chemistry, the formulation of electron-deficient bonds has been used for some time to describe nonclassical interactions between atoms. Traditional ligands in coordination chemistry such as amines and phosphanes bond to metal centers through their lone pair of electrons. Synergistic bonding effects dominate in the coordination of π-bonded ligands such as alkenes. In the mid-1980s the discovery of dihydrogen complexes having side-on coordination of H2 gave fresh impetus to transition metal chemistry as well as to the understanding of the interaction of σ-coordinating ligands with transition metals. In the meantime, transiton metal complexes can be obtained with a variety of σ-coordinated X-H fragments, and their mode of bonding can be understood by a common and quite general model. The chemistry of σ-bound silane ligands is particularly varied and well-investigated. These silane ligands enable the investigation of a large range of σ-coordinated metal complex fragments up to complete oxidative addition with cleavage of the Si? H bond and formation of silyl(hydrido) complexes, which has thus also widened our general understanding of the bonding of other σ-bound ligands. Whilst there is a large range of isolable and stable H2 and SiR4 complexes available, there are no such alkane analogues known at present. Only when the C? H bond is part of a ligand that is already directly bonded to the transition metal center will the resulting chelate effect stabilize this agostic C-H-M interaction. The complexation of SiH4, the simplest heavier homologue of CH4, was achieved recently. This is a further step towards the understanding of the factors which govern σ-complexation of ligands at transition metal centers.  相似文献   

14.
A theoretical study on two series of electron‐rich group 8 hydrides is carried out to evaluate involvement of the transition metal in dihydrogen bonding. To this end, the structural and electronic parameters are computed at the DFT/B3PW91 level for hydrogen‐bonded adducts of [(PP3)MH2] and [Cp*MH(dppe)] (M=Fe, Ru, Os; PP34‐P(CH2CH2PPh2)3, dppe= κ2‐Ph2PCH2CH2PPh2) with CF3CH2OH (TFE) as proton donor. The results are compared with those of adduct [Cp2NbH3] ? TFE featuring a “pure” dihydrogen bond, and classical hydrogen bonds in pyridine ? TFE and Me3N ? TFE. Deviation of the H ??? H? A fragment from linearity is shown to originate from the metal participation in dihydrogen bonding. The latter is confirmed by the electronic parameters obtained by NBO and AIM analysis. Considered together, orbital interaction energies and hydrogen bond ellipticity are salient indicators of this effect and allow the MH ??? HA interaction to be described as a bifurcate hydrogen bond. The impact of the M ??? HA interaction is shown to increase on descending the group, and this explains the experimental trends in mechanisms of proton‐transfer reactions via MH ??? HA intermediates. Strengthening of the M ??? H interaction in the case of electron‐rich 5d metal hydrides leads to direct proton transfer to the metal atom.  相似文献   

15.
Complexes with Macrocyclic Ligands. V Dinuclear Copper(II) Complexes with Chiral Macrocyclic Ligands of Schiff‐Base Type: Syntheses and Structures The synthesis and properties of four chiral, dinuclear, macrocyclic, cationic copper(II) complexes, [Cu2(Lm,n)]2+ ( 1 – 4 ), are described. The two symmetrical compounds [Cu2(L2,2)][ClO4]2 ( 1 and 2 ) were synthesized in a one‐step reaction from 2,6‐diformyl‐4‐tert.‐butylphenol, copper(II)‐perchlorate and the chiral diamine (1S,2S)‐1,2‐diphenylethylenediamine (synthesis of 1 ) and (1R,2R)‐1,2‐diaminocyclohexane (synthesis of 2 ), respectively. For the synthesis of the two unsymmetrical compounds [Cu2(LPh,n)][ClO4]2 ( 3 and 4 ) the mononuclear, neutral copper(II) complex [CuLPh] ( 5 ) [synthesized from 2,6‐diformyl‐4‐tert.‐butylphenol, copper(II)‐acetate and 1,2‐phenylenediamine] was reacted with (1R,2R)‐1,2‐diaminocyclohexane (synthesis of 3 ) and (S)‐1,1′‐binaphthyl‐2,2′‐diamine (synthesis of 4 ), respectively. The structures of the two unsymmetrical copper(II) compounds ( 3 and 4 ) were determined by X‐ray diffraction.  相似文献   

16.
A pair of novel azide‐bridged polynuclear copper(II) complexes, [CuL(μ1,1‐N3)]n ( 1 ) and [Cu4L2(CH3COO)21,1‐N3)4] ( 2 ) (L = 4‐chloro‐2‐[(2‐dimethylaminoethylimino)methyl]phenolate), have been obtained from the same Schiff base ligand and an identical synthetic procedure using anions of the metal salts as the only independent variable. Complex 1 was synthesized using copper(II) nitrate, while complex 2 was synthesized using the copper(II) acetate as the salt. Both of the complexes show novel supramolecular structures in their crystals as elucidated by X‐ray analyses. The polynuclear complex 1 contains [CuL(μ1,1‐N3)]n units as the building blocks, crystallizes in the Pbca space group. The tetra‐nuclear complex 2 contains [Cu4L2(CH3COO)21,1‐N3)4] units as the building blocks, crystallizes in the space group.  相似文献   

17.
Neutral and Cationic Ruthenium(II) Complexes with Trifunctional Phosphane Ligands Compounds of the type [RuCl2(RPX2)2] 4 – 7 (R = iPr, tBu; X = CH2CH2OMe, CH2CO2Me) were prepared by reacting RPX2 with either RuCl3 · 3H2O or [RuCl2(PPh3)3], respectively. In 4 – 7 the trifunctional phosphanes coordinate as bidentate ligands to the metal center through the phosphorus atom and the oxygen atom of a methoxy or carbonyl group. The lability of the Ru–O bond allows substitution reactions with CO, tert-butylisonitrile and phenylacetylene. The Ru–Cl bonds in 5 (R = tBu; X = CH2CH2OMe) can be cleaved upon treatment with one or two equiv of AgPF6 yielding mono- or dicationic derivatives. In these complexes the ligands are coordinated to the metal center through the phosphorus and both of the oxygen donor atoms. The reaction of the phosphinoesterenolate compound 17 with Ph2C=C=O leads to the insertion of two molecules of the ketene into the C–H bond of one of the five-membered metal-enolate rings to yield the “expanded” chelate complex 18 , the structure of which was determined by X-ray crystallography.  相似文献   

18.
Homoleptic carbonyl cations of the electron-rich metals in Groups 8 through 12 are the newest members of the large family of transition metal carbonyls. They can be distinguished from typical metal carbonyl complexes in several respects. Their synthesis entails carbonylation of metal salts in such superacids as fluorosulfuric acid and “magic acid” HSO3F? SbF5. Thermally stable salts with [Sb2F11]? as counterion are obtained with antimony pentafluoride as reaction medium. Both the [Sb2F11]? anion and superacid reaction media have previously found little application in the organometallic chemistry of d-block elements. Also unprecedented in metal carbonyl chemistry are the coordination geometries with coordination numbers 4 (square-planar coordination) and 2 (linear coordination) for the cation. Formal oxidation states of the metals, and the charges of the complex cations, extend from + 1 to +3: thus CO is largely σ-bonded to the metal, and the CO bond is strongly polarized. Minimal metal → CO π-backbonding and a positive partial charge on carbon are manifested in long M? C bonds, short C? O bonds, high frequencies for C? O stretching vibrations (up to 2300 cm?1), and small 13C NMR chemical shifts (up to δc, = 121). Prominent examples of these unusual homoleptic carbonyl cations, which were recently the subject of a Highlight in this journal, include the first carbonyl cation of a p-block metal [Hg(CO)2]2+, the first trivalent carbonyl cation [Ir(CO)6]3+, and the first multiply charged carbonyl cation of a 3d metal [Fe(CO)6]2+. In this overview we propose to (a) outline the historical origins of cationic metal carbonyls and their methods of synthesis; (b) present a summary of the general field of carbonyl cations, which has developed over a yery short period of time; (c) discuss the structural and spectroscopic characteritics of metal–CO bonding; (d) discuss the special significance associated with reaction media and the [Sb2F11]? anion; and (e) point to the most recent results and anticipated future developments.  相似文献   

19.
The pendant‐armed ligands L1 and L2 were synthesized by N‐alkylation of the four secondary amine groups of the macrocyclic precursor L using o‐nitrobenzylbromide (L1) and p‐nitrobenzylbromide (L2). Nitrates and perchlorates of CuII, NiII and CoII were used to synthesize the metal complexes of both ligands and the complexes were characterized by microanalysis, MS‐FAB, conductivity measurements, IR and UV‐Vis spectroscopy and magnetic studies. The crystal structures of L1, [CuL1](ClO4)2·CH3CN·H2O, [CuL2](ClO4)2·6CH3CN, [CuL2][Cu(NO3)4]·5CH3CN·0.5CH3OH and [NiL2](ClO4)2·3CH3CN·H2O were determined by single crystal X‐ray crystallography. These structural analysis reveal the free ligand L1, three mononuclear endomacrocyclic complexes {[CuL1](ClO4)2·CH3CN·H2O, [CuL2](ClO4)2·6CH3CN and [NiL2](ClO4)2·3CH3CN·H2O} and one binuclear complex {[CuL2][Cu(NO3)4]·5CH3CN·0.5CH3OH} in which one of the metals is in the macrocyclic framework and the other metal is outside the ligand cavity and coordinated to four nitrate ions.  相似文献   

20.
Monomeric Tripod–Zinc Thiolate Complexes Reaction of the pyrazolylborate-zinc complex TpCum,MeZn–OH with the corresponding thiols yielded the stable complexes TpCum,MeZn–SR ( 1 : R = Ph, 2 : R = CH2–Ph, 3 : R = CH2–CH2–Ph) which are further representatives of this class of compounds. Using the ligand tris(benzimidazolylmethyl)amine (BIMA), zinc perchlorate, and the corresponding sodium thiolates, the cationic complexes (BIMA)Zn–SR (R = Ph, CH2Ph) were obtained, which were isolated as [(BIMA)Zn–S–Ph] BPh4 ( 4 ) and [(BIMA)Zn–S–CH2Ph] ClO4 ( 5 ). A structure determination of 4 confirmed the pseudotetrahedral coordination geometry for this new type of compounds.  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