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 15 毫秒
1.
We previously reported on enantioselective aldol reactions of acetone and some aldehydes catalyzed by chiral Zn2+ complexes of L ‐prolyl‐pendant [12]aneN4 (L ‐ZnL1) and L ‐valyl‐pendant [12]aneN4 (L ‐ZnL2) in aqueous solution. Here, we report on the one‐pot chemoenzymatic synthesis of chiral 1,3‐diols in an aqueous solvent system at room temperature by a combination of enantioselective aldol reactions catalyzed by Zn2+ complexes of L ‐ and D ‐phenylalanyl‐pendant [12]aneN4 (L ‐ZnL3 and D ‐ZnL3) and the successive enantioselective reduction of the aldol products using oxidoreductases with the regeneration of the NADH (reduced form of nicotinamine adenine dinucleotide) cofactor. The findings indicate that all four stereoisomers of 1,3‐diols can be produced by appropriate selection of a chiral Zn2+‐complex and an oxidoreductase commercially available from the “Chiralscreen OH” kit.  相似文献   

2.
Aldolases are enzymes that catalyze stereospecific aldol reactions in a reversible manner. Naturally occurring aldolases include class I aldolases, which catalyze aldol reactions via enamine intermediates, and class II aldolases, in which Zn2+ enolates of substrates react with acceptor aldehydes. In this work, Zn2+ complexes of L ‐prolyl‐pendant[15]aneN5 (ZnL3), L ‐prolyl‐pendant[12]aneN4 (ZnL4), and L ‐valyl‐pendant[12]aneN4 (ZnL5) were designed and synthesized for use as chiral catalysts for enantioselective aldol reactions. The complexation constants for L3 to L5 with Zn2+ [logKs(ZnL)] were determined to be 14.1 (for ZnL3), 7.6 (for ZnL4), and 9.6 (for ZnL5), indicating that ZnL3 is more stable than ZnL4 and ZnL5. The deprotonation constants of Zn2+‐bound water [pKa(ZnL) values] for ZnL3, ZnL4, and ZnL5 were calculated to be 9.2 (for ZnL3), 8.2 (for ZnL4), and 8.6 (for ZnL5), suggesting that the Zn2+ ions in ZnL3 is a less acidic Lewis acid than in ZnL4 and ZnL5. These values also indicated that the amino groups on the side chains weakly coordinate to Zn2+. We carried out aldol reactions between acetone and 2‐chlorobenzaldehyde and other aldehydes in the presence of catalytic amounts of the chiral Zn2+ complexes in acetone/H2O at 25 and 37 °C. Whereas ZnL3 yielded the aldol product in 43 % yield and 1 % ee (R), ZnL4 and ZnL5 afforded good chemical yields and high enantioselectivities of up to 89 % ee (R). UV titrations of proline and ZnL4 with acetylacetone (acac) in DMSO/H2O (1:2) indicate that ZnL4 facilitates the formation of the ZnL4 ? (acac)? complex (Kapp=2.1×102 M ?1), whereas L ‐proline forms a Schiff base with acac with a very small equilibrium constant. These results suggest that the amino acid components and the Zn2+ ions in ZnL4 and ZnL5 function in a cooperative manner to generate the Zn2+‐enolate of acetone, thus permitting efficient enantioselective C? C bond formation with aldehydes.  相似文献   

3.
The synthesis of a new ligand (L1) containing two 1,4,7‐triazacyclononane ([9]aneN3) moieties linked by a 4,5‐dimethylenacridine unit is reported. The binding and fluorescence sensing properties toward Cu2+, Zn2+, Cd2+, and Pb2+ of L1 and receptor L2, composed of two [9]aneN3 macrocycles bridged by a 6,6′′‐dimethylen‐2,2′:6′,2′′‐terpyridine unit, have been studied by coupling potentiometric, UV/Vis absorption, and emission measurements in aqueous media. Both receptors can selectively detect Zn2+ thanks to fluorescence emission enhancement upon metal binding. The analysis of the binding and sensing properties of the Zn2+ complexes toward inorganic anions revealed that the dinuclear Zn2+ complex of L1 selectively binds and senses the triphosphate anion (TP), whereas the mononuclear Zn2+ complex of L2 displays selective recognition of diphosphate (DP). Binding of TP or DP induces emission quenching of the Zn2+ complexes with L1 and L2, respectively. These results are exploited to discuss the role played by pH, number of coordinated metal cations, and binding ability of the bridging units in metal and/or anion coordination and sensing.  相似文献   

4.
Two complexes, cis‐[MnL2(NCS)2] ( 1 ) and cis‐[ZnL2(NCS)2] ( 2 ) with asymmetrical substituted triazole ligands [L = 3,4‐dimethyl‐5‐(2‐pyridyl)‐1,2,4‐triazole], were synthesized and characterized by elemental analysis, UV/Vis and FT‐IR spectroscopy as well as thermogravimetric analyses (TGA), powder XRD, and single‐crystal X‐ray diffraction. In the complexes, each L molecule adopts a chelating bidentate mode by the nitrogen atoms of pyridyl and triazole. Both complexes have a similar distorted octahedral [MN6] core (M = Mn2+ and Zn2+) with two NCS ions in the cis position.  相似文献   

5.
Two new flexible extended dialdehydes (H2hpdd and H2pdd) with different functional pendant arms (? CH2CH2PhOH and ? CH2CH2Ph) have been synthesized and reacted with 1,2‐bis(2‐aminoethoxy)ethane to prepare Schiff‐base macrocyclic complexes in the presence of a ZnII‐ion template. As a result, two preorganized dinuclear ZnII intermediates ( 1 and 2 ), as well as two 42‐membered folded [2+2] macrocyclic dinuclear ZnII complexes ( 3 and 4 ), were produced. The central zinc ions in compounds 1 – 4 showed distinguishable coordination patterns with the dialdehydes and the [2+2] macrocyclic ligands, in which a subtle pH‐adjustment function of the two pendant arms (with or without the phenolic hydroxy group) was believed to play a vital role. Furthermore, cation‐ and anion‐recognition experiments for complexes 3 and 4 revealed that they could selectively recognize acetate ions by the formation of 1:1 stoichiometric complexes, as verified by changes in their UV/Vis and MS (ESI) spectra and even by the naked eye.  相似文献   

6.
This work demonstrates a selection criteria that determines whether molecular assembly occurs through a one‐step or stepwise manner in ligand‐bridged dinuclear zinc(II) (Zn2+) complex formation, which is associated with the π stacking of building blocks. The building blocks of carbazole ligands ( L1 and L4 ) that contain two imidazole moieties at the 3,6‐positions form 4:2 complexes (i.e., [ L ]4?(Zn2+)2) at a molar ratio of 0.50 ([Zn2+]/[ L ]0=0.50), thereby providing π stacking between the carbazole ligands. At the molar ratio of 0.67 ([Zn2+]/[ L ]0=0.67), the 4:2 complexes change to 3:2 complexes (i.e., [ L ]3?(Zn2+)2) with no π‐stacked carbazole unit. In contrast, when the imidazole groups in L1 are replaced with benzoimidazole groups ( L3 ), L3 also yields the 4:2 complex [( L3 )4?(Zn2+)2] at a molar ratio of 0.50. However, there is no structural transition from ( L3 )4?(Zn2+)2 to other complex species above a molar ratio of 0.50. Similarly, when two imidazole groups are introduced into the carbazole ring at 2,7‐positions ( L5 ), L5 also gives the 4:2 complex [( L5 )4?(Zn2+)2] that shows no structural transition to other complex species at a higher molar ratio.  相似文献   

7.
Seven discrete sugar‐pendant diamines were complexed to the {M(CO)3}+ (99mTc/Re) core: 1,3‐diamino‐2‐propyl β‐D ‐glucopyranoside ( L 1 ), 1,3‐diamino‐2‐propyl β‐D ‐xylopyranoside ( L 2 ), 1,3‐diamino‐2‐propyl α‐D ‐mannopyranoside ( L 3 ), 1,3‐diamino‐2‐propyl α‐D ‐galactopyranoside ( L 4 ), 1,3‐diamino‐2‐propyl β‐D ‐galactopyranoside ( L 5 ), 1,3‐diamino‐2‐propyl β‐(α‐D ‐glucopyranosyl‐(1,4)‐D ‐glucopyranoside) ( L 6 ), and bis(aminomethyl)bis[(β‐D ‐glucopyranosyloxy)methyl]methane ( L 7 ). The Re complexes [Re( L 1 – L 7 )(Br)(CO)3] were characterized by 1H and 13C 1D/2D NMR spectroscopy which confirmed the pendant nature of the carbohydrate moieties in solution. Additional characterization was provided by IR spectroscopy, elemental analysis, and mass spectrometry. Two analogues, [Re( L 2 )(CO)3Br] and [Re( L 3 )(CO)3Br], were characterized in the solid state by X‐ray crystallography and represent the first reported structures of Re organometallic carbohydrate compounds. Conductivity measurements in H2O established that the complexes exist as [Re( L 1 – L 7 )(H2O)(CO)3]Br in aqueous conditions. Radiolabelling of L 1 – L 7 with [99mTc(H2O)3(CO)3]+ afforded in high yield compounds of identical character to the Re analogues. The radiolabelled compounds were determined to exhibit high in vitro stability towards ligand exchange in the presence of an excess of either cysteine or histidine over a 24 h period.  相似文献   

8.
Previous studies into the dissociation of [CuII(dien)peptide] . 2+ ions (dien = diethylenetriamine) have shown that NH‐containing auxiliary ligands do not favor the formation of [peptide] . + species; instead, they promote proton‐transfer reactions, especially for peptides containing basic amino residues. Formation of radical cationic tripeptides of the form GGX . + [GGX = glycylglycyl(residue X)] becomes feasible upon substituting the open‐chain tridentate ligand dien with its analogous cyclic ligand, 1,4,7‐triazacyclononane (9‐aneN3); i.e., from [CuII(9‐aneN3)GGX] . 2+ ions. Similar enhancements occur when using 1,4,7,10‐tetraoxacyclododecane (12‐crown‐4) in place of its open‐chain analog, 2,5,8,11‐tetraoxadecane (triglyme). We have demonstrated that a sterically encumbered auxiliary macrocyclic ligand within [CuII(L)GGX] . 2+ complex ions [where L = 9‐aneN3 or 12‐crown‐4] facilitates the formation of radical cationic peptides through gas‐phase fragmentation. We verified our experimental observations by examining the reactivities of a series of 19 tripeptides of the type GGX that differ only in the identity of their C‐terminal residue. The energy of the electron‐transfer reaction correlates well with the bond‐dissociation energy of the peptide–Cu(II) interaction; the presence of a constrained macrocyclic ligand weakens metal–peptide chelation through steric repulsion between the ligand and the peptide, and this situation may lead to more favorable radical cationic peptide formation. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

9.
A series of five l ‐di‐p‐toluoyl‐tartaric acid (l ‐DTTA) lanthanide coordination polymers, namely {[Ln4K4 L6(H2O)x]?yH2O}n, [Ln=Dy ( 1 ), x=24, y=12; Ln=Ho ( 2 ), x=23, y=12; Ln=Er ( 3 ), x=24, y=12; Ln=Yb ( 4 ), x=24, y=11; Ln=Lu ( 5 ), x=24, y=12] have been isolated by simple reactions of H2L (H2L= L ‐DTTA) with LnCl3?6 H2O at ambient temperature. X‐ray crystallographic analysis reveals that complexes 1 – 5 feature two‐dimensional (2D) network structures in which the Ln3+ ions are bridged by carboxylate groups of ligands in two unique coordinated modes. Luminescent spectra demonstrate that complex 1 realizes single‐component white‐light emission, while complexes 2 – 4 exhibit a characteristic near‐infrared (NIR) luminescence in the solid state at room temperature.  相似文献   

10.
Summary The formation constants of species formed in the systems H+-Zn2+-cysteine and H+-Zn2+-cystine have been determined in aqueous solution at 37° and I = 0.15 mot dm–3 (NaClO4), using the pH-metric method. The existence of the following species [ZnL], [ZnL2], [ZnL2H] and [Zn2L3] (2.3 pH 7.7) was proved for the Zn2+-cysteine system, whereas for the Zn2+-cystine [Zn2L] (5.3 pH 6.4) was the only species found. In the Zn2+-cystine system the pH range was severely restricted because of precipitation occurring at pH > 6.4. A new experimental and numerical approach was employed in order to implement the possibility of rigorously selecting the species present in each system. The results have been compared with data previously reported on the same systems, considering in particular the different sets of species found in the various works.  相似文献   

11.
Coordination studies on Zn(II) complexes of 1,3,5‐tri(2,5‐diazahexyl)benzene (L) show that by comparison with the non‐deprotonation of complex ZnL in a 1:1 system, the three‐dimensional complexiaton decreases the pKa of the Zn‐bound water molecule, that is, pKa = 7.47 for trinulclear complex Zn3L in a 3:1 metal–ligand ratio. These two types of zinc(II) complexes have been examined as catalysts for the hydrolysis of 4‐nitrophenyl acetate (NA) in 10% (v/v) CH3CN at 298 K, I = 0.10 mol dm?3 KNO3 at pH range 6.5–8.2 and 8.5–10, respectively. Kinetic studies show that the second‐order rate constants of NA‐hydrolysis catalyzed by complexes ZnL, Zn3L, and Zn3LH?1 are 0.021, 0.0082, and 0.342 mol?1 dm3 s?1, respectively. In all the cases, the pH‐dependent observed first‐order rate constant, kobs, shows sigmoidal pH–rate profile. The 1:1 complex ZnL–H2O undergoes NA hydrolysis by direct rate‐determining hydrolysis to produce 4‐nitrophenol(ate) (NP?) and ZnL(OOCCH3); while in the 3:1 system the oxygen atom of acetic group forms a H‐bond with the Zn(II)‐bound water of the second branch of tripod indicating that the polynuclear centers are associated and bi‐functional. © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 36: 41–48 2004  相似文献   

12.
Copper(I) halides with triphenyl phosphine and imidaozlidine‐2‐thiones (L ‐NMe, L ‐NEt, and L ‐NPh) in acetonitrile/methanol (or dichloromethane) yielded copper(I) mixed‐ligand complexes: mononuclear, namely, [CuCl(κ1‐S‐L ‐NMe)(PPh3)2] ( 1 ), [CuBr(κ1‐S‐L ‐NMe)(PPh3)2] ( 2 ), [CuBr(κ1‐S‐L ‐NEt)(PPh3)2] ( 5 ), [CuI(κ1‐S‐L ‐NEt)(PPh3)2] ( 6 ), [CuCl(κ1‐S‐L ‐NPh)(PPh3)2] ( 7 ), and [CuBr(κ1‐S‐L ‐NPh)(PPh3)2] ( 8 ), and dinuclear, [Cu21‐I)2(μ‐S‐L ‐NMe)2(PPh3)2] ( 3 ) and [Cu2(μ‐Cl)21‐S‐L ‐NEt)2(PPh3)2] ( 4 ). All complexes were characterized with analytical data, IR and NMR spectroscopy, and X‐ray crystallography. Complexes 2 – 4 , 7 , and 8 each formed crystals in the triclinic system with P$\bar{1}$ space group, whereas complexes 1 , 5 , and 6 crystallized in the monoclinic crystal system with space groups P21/c, C2/c, and P21/n, respectively. Complex 2 has shown two independent molecules, [(CuBr(κ1‐S‐L ‐NMe)(PPh3)2] and [CuBr(PPh3)2] in the unit cell. For X = Cl, the thio‐ligand bonded to metal as terminal in complex 4 , whereas for X = I it is sulfur‐bridged in complex 3 .  相似文献   

13.
The Zn complexes bis(acetylacetonato‐κ2O,O′)bis{4′‐[4‐(methylsulfanyl)phenyl]‐4,2′:6′,4′′‐terpyridine‐κN1}zinc(II), [Zn(C5H7O2)2(C22H17N3S)2], (I), and {μ‐4′‐[4‐(methylsulfanyl)phenyl]‐4,2′:6′,4′′‐terpyridine‐κ2N1:N1′′}bis[bis(acetylacetonato‐κ2O,O′)zinc(II)], [Zn2(C5H7O2)4(C22H17N3S)], (II), are discrete entities with different nuclearities. Compound (I) consists of two centrosymmetrically related monodentate 4′‐[4‐(methylsulfanyl)phenyl]‐4,2′:6′,4′′‐terpyridine (L1) ligands binding to one ZnII atom sitting on an inversion centre and two centrosymmetrically related chelating acetylacetonate (acac) groups which bind via carbonyl O‐atom donors, giving an N2O4 octahedral environment for ZnII. Compound (II), however, consists of a bis‐monodentate L1 ligand bridging two ZnII atoms from two different Zn(acac)2 fragments. Intra‐ and intermolecular interactions are weak, mainly of the C—H...π and π–π types, mediating similar layered structures. In contrast to related structures in the literature, sulfur‐mediated nonbonding interactions in (II) do not seem to have any significant influence on the supramolecular structure.  相似文献   

14.
Two trinuclear CoII and ZnII complexes, [(CoL)2(OAc)2Co] and [(ZnL)2(OAc)2Zn], with an asymmetric Salen‐type bisoxime ligand [H2L = 4‐(N,N‐diethylamine)‐2,2′‐[ethylenediyldioxybis(nitrilomethylidyne)]diphenol] were synthesized and characterized by elemental analyses, IR, UV/Vis, and fluorescent spectroscopy. The crystal structures of the CoII and ZnII complexes were determined by single‐crystal X‐ray diffraction methods. The CoII atom is pentacoodinated by N2O2 donor atoms from the (L)2– unit and one oxygen atom from the coordinated acetate ion, resulting in a trigonal bipyramid arrangement. With the help of intermolecular hydrogen bonding C–H ··· O and C–H ··· π interactions, a self‐assembled continual zigzag chain‐like supramolecular structure is formed. The ZnII atom is pentacoodinated by N2O2 donor atoms from the (L)2– unit and one oxygen atom from the coordinated acetate ion, resulting in an almost regular trigonal bipyramid arrangement. A self‐assembled continual 1D supramolecular chain‐like structure is formed by intermolecular hydrogen bonding C–H ··· O and C–H ··· π interactions. Additionally, the photophysical properties of the CoII and ZnII complexes were discussed.  相似文献   

15.
A novel tridentate anilido‐aldimine ligand, [o‐C6H4(NHAr)? HC?NCH2CH2NMe2] (Ar = 2,6‐iPr2C6H3, L ‐H, 1 ), has been prepared by the condensation of N, N‐dimethylethylenediamine with one molar equivalent of 2‐fluoro‐benzaldehyde in hexane, followed by the addition of the lithium salt of diisopropylaniline in THF. Magnesium (Mg) and zinc (Zn) complexes supported by the tridentate anilido‐aldimine ligand have been synthesized and structurally characterized. Reaction of L ‐H ( 1 ) with an equivalent amount of MgnBu2 or ZnEt2 produces the monomeric complex [ L MgnBu] ( 2 ) or [ L ZnEt] ( 3 ), respectively. Experimental results show that complexes 2 and 3 are efficient catalysts for ring‐opening polymerization of ε‐caprolactone (CL) and L ‐lactide (LA) in the presence of benzyl alcohol and catalyze the polymerization of ε‐CL and L ‐LA in a controlled fashion yielding polymers with a narrow polydispersity index. In both polymerizations, the activity of Mg complex 2 is higher than that of Zn complex 3 , which is probably due to the higher Lewis acidity and better oxophilic nature of Mg2+ metal. © 2009 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 4927–4936, 2009  相似文献   

16.
Ligand L (4‐(7‐nitrobenzo[1,2,5]oxadiazole‐4‐yl)‐1,7‐dimethyl‐1,4,7,10‐tetra‐azacyclododecane) is a versatile fluorescent sensor useful for CuII, ZnII and CdII metal detection, as a building block of fluorescent metallo‐receptor for halide detection, and as an organelle marker inside live cells. Ligand L undergoes a chelation‐enhanced fluorescence (CHEF) effect upon metal coordination in acetonitrile solution. In all three complexes investigated the metal cation is coordinatively unsaturated; thus, it can bind secondary ligands as anionic species. The crystal structure of [Zn L Cl](ClO4) is discussed. CuII and ZnII complexes are quenched upon halide interaction, whereas the [Cd L ]2+ species behaves as an OFF–ON sensor for halide anions in acetonitrile solution. The mechanism of the fluorescence response in the presence of the anion depends on the nature of the metal ion employed and has been studied by spectroscopic methods, such as NMR spectroscopy, UV/Vis and fluorescence techniques and by computational methods. Subcellular localization experiments performed on HeLa cells show that L mainly localizes in spot‐like structures in a polarized portion of the cytosol that is occupied by the Golgi apparatus to give a green fluorescence signal.  相似文献   

17.
The sequential reaction of a multisite coordinating compartmental ligand [2‐(2‐hydroxy‐3‐(hydroxymethyl)‐5‐methylbenzylideneamino)‐2‐methylpropane‐1,3‐diol] (LH4) with appropriate lanthanide salts followed by the addition of [Mg(NO3)2] ? 6 H2O or [Zn(NO3)2] ? 6 H2O in a 4:1:2 stoichiometric ratio in the presence of triethylamine affords a series of isostructural heterometallic trinuclear complexes containing [Mg2Ln]3+ (Ln=Dy, Gd, and Tb) and [Zn2Ln]3+ (Ln=Dy, Gd, and Tb) cores. The formation of these complexes is demonstrated by X‐ray crystallography as well as ESI‐MS spectra. All complexes are isostructural possessing a linear trimetallic core with a central lanthanide ion. The comprehensive studies discussed involve the synthesis, structure, magnetism, and photophysical properties on this family of trinuclear [Mg2Ln]3+ and [Zn2Ln]3+ heterometallic complexes. [Mg2Dy]3+ and [Zn2Dy]3+ show slow relaxation of the magnetization below 12 K under zero applied direct current (dc) field, but without reaching a neat maximum, which is due to the overlapping with a faster quantum tunneling relaxation mediated through dipole–dipole and hyperfine interactions. Under a small applied dc field of 1000 Oe, the quantum tunneling is almost suppressed and temperature and frequency dependent peaks are observed, thus confirming the single‐molecule magnet behavior of complexes [Mg2Dy]3+ and [Zn2Dy]3+.  相似文献   

18.
Four novel mononuclear ruthenium(II) complexes [Ru(dmb)2L]2+ [dmb = 4,4′‐dimethyl‐2,2′‐bipyridine, L = imidazo‐[4,5‐f][1,10]phenanthroline (IP), 2‐phenylimidazo‐[4,5‐f][1,10]phenanthroline (PIP), 2‐(4′‐hydroxyphenyl)imidazo‐[4,5‐f] [1,10] phenanthroline (HOP), 2‐(4′‐dimethylaminophenyl) imidazo‐[4, 5‐f] [1,10] phenanthroline (DMNP)] were synthesized and characterized by ES‐MS, 1H NMR, UV‐vis and electrochemistry. The nonlinear optical properties of the ruthenium(II) complexes were investigated by Z‐scan techniques with 12 ns laser pulse at 540 nm, and all of them exhibit both nonlinear optical (NLO) absorption and self‐defocusing effect. The corresponding effective NLO susceptibility |x3| of the complexes is in the range of 2.68 × 10?12‐4.57 × 10?12 esu.  相似文献   

19.
Postsynthetic installation of lanthanide cubanes into a metallosupramolecular framework via a single‐crystal‐to‐single‐crystal (SCSC) transformation is presented. Soaking single crystals of K6[Rh4Zn4O(l ‐cys)12] (K6[ 1 ]; l ‐H2cys=l ‐cysteine) in a water/ethanol solution containing Ln(OAc)3 (Ln3+=lanthanide ion) results in the exchange of K+ by Ln3+ with retention of the single crystallinity, producing Ln2[ 1 ] ( 2Ln ) and Ln0.33[Ln4(OH)4(OAc)3(H2O)7][ 1 ] ( 3Ln ) for early and late lanthanides, respectively. While the Ln3+ ions in 2Ln exist as disordered aqua species, those in 3Ln form ordered hydroxide‐bridged cubane clusters that connect [ 1 ]6? anions in a 3D metal‐organic framework through coordination bonds with carboxylate groups. This study shows the utility of an anionic metallosupramolecular framework with carboxylate groups for the creation of a series of metal cubanes that have great potential for various applications, such as magnetic materials and heterogeneous catalysts.  相似文献   

20.
2D 1H,89Y heteronuclear shift correlation through scalar coupling has been applied to the chemical‐shift determination of a set of yttrium complexes with various nuclearities. This method allowed the determination of 89Y NMR data in a short period of time. Multinuclear NMR spectroscopy as function of temperature, PGSE NMR‐diffusion experiments, heteronuclear NOE measurements, and X‐ray crystallography were applied to determine the structures of [Y5(OH)5(L ‐Val)4(Ph2acac)6] ( 1 ) (Ph2acac=dibenzoylmethanide, L ‐Val=L ‐valine), [Y( 2 )(OTf)3] ( 3 ), and [Y2( 4 )(OTf)5] ( 5 ) ( 2 : [(S)P{N(Me)N?C(H)Py}3], 4 : [B{N(Me)N?C(H)Py}4]?) in solution and in the solid state. The structures found in the solid state are retained in solution, where averaged structures were observed. NMR diffusion measurements helped us to understand the nuclearity of compounds 3 and 5 in solution. 1H,19F HOESY and 19F,19F EXSY data revealed that the anions are specifically located in particular regions of space, which nicely correlated with the geometries found in the X‐ray structures.  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