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 593 毫秒
1.
The ability of the transition metal complex M(salen)* (M = Ni, Cu) to form Lewis acid-base adducts with lead(II) salts has been explored. The new complexes Pb(Hsal)(2)(Cu(salen*))(2) (1), [Pb(NO(3))(Cu(salen*))(2)](NO(3)) (2), Pb(OAc)(2)(Cu(salen*)) (3), and [Pb(OAc)(Ni(salen*)(2)](OAc) (4) (Hsal = O(2)CC(6)H(4)-2-OH, salen* = bis(3-methoxy)salicylideneimine) have been synthesized and characterized spectroscopically and by single-crystal X-ray diffraction. The coordination environment of the lead in the heterobimetallic complex is sensitive both to the initial lead salt and to the transition metal salen* complex that is employed in the synthesis. As a result, we have been able to access both 2:1 and 1:1 adducts by varying either the lead salt or the transition metal in the heterobimetallic coordination complex. In all cases, the salen* complex is associated with the lead center via dative interactions of the phenolic oxygen atoms. The relationship between the coordination requirements of the lead and the chemical nature of the anion is examined. In compound 1, the Pb(2+) ion is chelated by two Cu(salen*) moieties, and both salicylate ligands remain attached to the lead center and bridge to the Cu(2+) ions. The two Cu(salen*) groups are roughly parallel and opposed to each other as required by crystallographic inversion symmetry at lead. In contrast, the two Cu(salen*) groups present in 2 and 4 attached to the lead ion show considerable overlap. Furthermore, only one nitrate ion in 2 and one acetate ion in 4 remain bonded to the lead center. Compound 3 is unique in that only one Cu(salen*) group can bind to lead. Here, both acetate ligands remain attached, although one is chelating bidentate and the other is monodentate.  相似文献   

2.
水热法合成了配位聚合物[La(EBA)3(EBAH)(H2O)]n(EBA=4-乙基苯甲酸根,EBAH=4-乙基苯甲酸),并通过X-射线衍射单晶结构分析、红外光谱、紫外光谱、荧光光谱以及热重分析对配合物进行了结构和性质研究。配合物属三斜晶系,P1空间群。该配合物具有一维链状结构。La3+离子与9个O原子配位,其中4个O原子来自4个双齿桥联的4-乙基苯甲酸根,3个O原子来自2个螯合-桥联的4-乙基苯甲酸根,1个O原子来自1个中性的4-乙基苯甲酸,1个O原子来自配位水分子。La3+离子处于九配位扭变的三帽三角棱柱构型中。相邻La3+离子通过双齿桥联或螯合桥联的4-乙基苯甲酸根联结成一维链状结构。存在于分子内的氢键使一维链状结构更加稳定。由于相邻一维链的苯环间存在弱的π…π堆积作用,使分子沿着a轴堆积形成二维层状结构。同时,标题配合物固体具有光致发光特性,蓝光区的较强发射归于配体的π→π*电子跃迁。  相似文献   

3.
The tetraphenylborate salt of the decamethyl titanocene cation, [Cp*2Ti][BPh4] (1, Cp* = C5Me5), was prepared by reaction of Cp*2TiH with [Cp2Fe][BPh4] and by reaction of Cp*2TiMe with [PhNMe2H][BPh4]. The crystal structure of 1 shows that the Cp*2Ti cation has a bent metallocene structure with agostic interactions with the metal center of two adjacent methyl groups on one of the Cp* ligands. Compound 1 reacts readily with THF to give the adduct [Cp*2Ti(THF)][BPh4] (2). In fluorobenzene, 1 forms the eta1-fluorobenzene adduct [Cp*2Ti(eta1-FC6H5)][BPh4] (3), which was structurally characterized. In contrast to the thermal stability of 3, addition of alpha,alpha,alpha-trifluorotoluene to either 1 or 2 results in C-F activation to give Cp*2TiF2 and PhCF2CF2Ph as the main products. This reactivity toward benzylic C-F bonds is also reflected in the reactivity toward the fluorinated borate anions [B(C6F5)4]- and {B(3,5-(CF3)2C6H3]4}-: reaction of Cp*2TiMe with their [PhNMe2H]+ salts results in a stable complex for the former anion, whereas rapid C-F activation is observed for the latter.  相似文献   

4.
铅(II)-氨基多羧酸配合物的合成与结构研究   总被引:1,自引:0,他引:1  
合成了铅(II)与乙二胺四乙酸(EDTA),N-羟乙基乙二胺三乙酸(HEDTA)和 (乙酸-N-羟乙基酯)乙二胺三乙酸(AHEDTA, A = 乙酸-N-羟乙基酯)的配合物,C10H20K2N2O12Pb (K2[Pb(EDTA)]4H2O), C10H22K2N2O11Pb (K2[Pb(HEDTA)]4H2O)和C12H23KN2O11Pb (K[Pb- (AHEDTA)]3H2O), 并测定了K2[Pb(EDTA)]4H2O晶体结构和分子结构。具体测定结果如下:单斜晶系,C2/c空间群,a = 24.59(2),b = 11.79(1),c = 14.08(2) 牛琤 = 108.15(2),V = 3876.0(7)) 3,Z = 8,Mr = 654.65,Dc = 2.213 g/cm3,m = 9.196 mm-1和F(000) = 2480。最终偏差因子分别为R = 0.0458,wR = 0.0640 (对3372 (I > 2.0s(I))可观测衍射点)。在K2[Pb(EDTA)]4H2O中,配合物离子[Pb(EDTA)]2-具有六配位的非标准三棱柱体结构,EDTA作为六齿配体提供4个O原子和2个N原子与中心金属离子Pb2+形成配键。  相似文献   

5.
To investigate the potential role of Sm-Ph species as intermediates in the samarium-catalyzed redistribution of PhSiH3 to Ph2SiH2 and SiH4, the samarium phenyl complex [Cp*2SmPh]2 (1) was prepared by oxidation of Cp2*Sm (2) with HgPh2. Compound 1 thermally decomposes to yield benzene and the phenylene-bridged disamarium complex Cp*2Sm(mu-1,4-C6H4)SmCp*2 (3). This decomposition reaction appears to proceed through dissociation of 1 into monomeric Cp*2SmPh species which then react via unimolecular and bimolecular pathways, involving rate-limiting Cp* metalation and direct C-H activation, respectively. The observed rate law for this process is of the form: rate = k1[1] + k2[1]2. Complex 1 efficiently transfers its phenyl group to PhSiH3, with formation of Ph2SiH2 and [Cp*2Sm(mu-H)]2 (4). Quantitative Si-C bond cleavage of C6F5SiH3 is effected by the samarium hydride complex 4, yielding silane and [Cp*2Sm(mu-C6F5)]2 (5). In contrast, Si-H activation takes place upon reaction of 4 with o-MeOC6H4SiH3, affording the samarium silyl species [structure: see text] Cp*2SmSiH2(o-MeOC6H4) (7). Complex 7 rapidly decomposes to [Cp*2Sm(mu-o-MeOC6H4)]2 (6) and other samarium-containing products. Compounds 5 and 6 were prepared independently by oxidation of 2 with Hg(C6F5)2 and Hg(o-MeOC6H4)2, respectively. The mechanism of samarium-mediated redistribution at silicon, and chemoselectivity in sigma-bond metathesis reactions, are discussed.  相似文献   

6.
In an attempt to assess the potential role of the hydroxyl radical in the atmospheric degradation of sulfuric acid, the hydrogen transfer between H2SO4 and HO* in the gas phase has been investigated by means of DFT and quantum-mechanical electronic-structure calculations, as well as classical transition state theory computations. The first step of the H2SO4 + HO* reaction is the barrierless formation of a prereactive hydrogen-bonded complex (Cr1) lying 8.1 kcal mol(-1) below the sum of the (298 K) enthalpies of the reactants. After forming Cr1, a single hydrogen transfer from H2SO4 to HO* and a degenerate double hydrogen-exchange between H2SO4 and HO* may occur. The single hydrogen transfer, yielding HSO4* and H2O, can take place through three different transition structures, the two lowest energy ones (TS1 and TS2) corresponding to a proton-coupled electron-transfer mechanism, whereas the higher energy one (TS3) is associated with a hydrogen atom transfer mechanism. The double hydrogen-exchange, affording products identical to reactants, takes place through a transition structure (TS4) involving a double proton-transfer mechanism and is predicted to be the dominant pathway. A rate constant of 1.50 x 10(-14) cm(3) molecule(-1) s(-1) at 298 K is obtained for the overall reaction H2SO4 + HO*. The single hydrogen transfer through TS1, TS2, and TS3 contributes to the overall rate constant at 298 K with a 43.4%. It is concluded that the single hydrogen transfer from H2SO4 to HO* yielding HSO4* and H2O might well be a significant sink for gaseous sulfuric acid in the atmosphere.  相似文献   

7.
On interacting 1,10-phenanthroline (phen) with tetrabutylammonium tetrafluoroborate (Bu4N*BF4) two novel compounds have been obtained. These are the 1:1 adduct (phen)*(Bu4N*BF4) and the complex Pb(phen)(Bu4N*BF4)3(NO3)2. The synthesized substances have been characterized by UV, IR and Raman spectroscopy as well as by 1H- and 19F-NMR spectroscopy.  相似文献   

8.
An excimer-based, binuclear, on-off switchable calix[4]crown chemosensor   总被引:3,自引:0,他引:3  
A new fluorescent chemosensor with two different types of cation binding sites on the lower rims of a 1,3-alternate calix[4]arene (1) is synthesized. Two pyrene moieties linked to a cation recognition unit composed of two amide groups form a strong excimer in solution. For 1, the excimer fluorescence is quenched by Pb2+, but revived by addition of K+ to the Pb2+ ligand complex. Thus, metal ion exchange produces an on-off switchable, fluorescent chemosensor. Computational results show that the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbitals (LUMO) of the two pyrene moieties interact under UV irradiation of 1 and its K+ complex, while such HOMO-LUMO interactions are absent in the Pb2+ complex.  相似文献   

9.
孔德轮  高保娇  李刚 《物理化学学报》2006,22(11):1399-1403
采用粘度法确定了聚4-乙烯基吡啶(P4VP)在乙醇/水混合溶剂中的临界交迭浓度c*, 分别在稀溶液与亚浓溶液浓度范围内, 采用光谱法与电导滴定法研究了P4VP与Cu(II)离子的配合过程及配合物的结构, 通过红外光谱(FTIR)对配合物的化学结构进行了表征, 并用差示扫描量热法(DSC)测定了配合物的热性能. 结果表明, 对于相对分子质量为1.06×105的P4VP, 其c*为15 mmol•L−1(按P4VP中的链节量计算). 在稀溶液中P4VP与Cu(II)离子形成可溶性的分子内配合物, 表观配位数为9~10;在亚浓溶液中, P4VP与Cu(II)离子发生分子间配合作用, 由于配位交联, 形成不溶性的配合物P4VP-Cu(II), 配位数为3. P4VP与Cu(II)离子形成配合物后, 玻璃化温度明显提高.  相似文献   

10.
A hydrogen bond formed between the superoxide anion and the ammonium ion (NH4+) accelerates electron transfer from the C60 radical anion to oxygen significantly, whereas the tetra-n-butylammonium ion has no ability to form a hydrogen bond with the superoxidie anion, exhibiting no acceleration of the electron-transfer reduction of oxygen. The second-order rate constant of electron transfer from C60*- to O2 increases linearly with increasing concentration of NH4+. This indicates that O2*- produced in the electron transfer from C60 to O2 is stabilized by 1:1 complex formation between O2*- and NH4+. The 1:1 complex formed between O2*- and NH4+ was detected by ESR. The binding of O2*- with NH4+ results in a positive shift of the reduction potential of O2 with increasing concentration of NH4+, leading to the acceleration of electron transfer from C60*- to O2.  相似文献   

11.
Fan SR  Zhu LG 《Inorganic chemistry》2007,46(16):6785-6793
Four structurally diverse complexes, {[Pb(Hssal)(2,2'-bipy)](4,4'-bipy)0.5}n (1), [Pb2(Hssal)2(2,2'-bipy)2(4,4'-bipy)(H2O)2] (2), [Pb(Hssal)(phen)(4,4'-bipy)0.5]n (3), and [Pb(Hssal)(2,2'-bipy)(bpe)0.5]n (4), have been synthesized and characterized by elemental analyses, IR, thermogravimetric analyses, fluorescent spectra, and single-crystal X-ray analyses, where Hssal2- is doubly deprotonated 5-sulfosalicylate, 2,2'-bipy is 2,2'-bipyridine, phen is 1,10-phenanthroline, 4,4'-bipy is 4,4'-bipyridine, and bpe is trans-1,2-bis(4-pyridyl)ethylene. The structure of complex 1 possesses a one-dimensional ladderlike chain with guest 4,4'-bipy molecules, while the molecular structure of complex 2 is a dimeric species with a coordinating 4,4'-bipy ligand. Complex 3 consists of a one-dimensional ladderlike chain with monodentate 4,4'-bipyridine but somewhat different from that of complex 1. Complex 4 is a two-dimensional layer structure. In 1-4, all 5-sulfosalicylates are doubly deprotonated, and all carboxylate groups of Hssal2- chelate to PbII ions; however, the coordination modes of sulfonyl groups are different: syn-syn bridging in 1, noncoordinating in 2, syn-skew bridging in 3, and one-atom bridging in 4. The noncoordinating mode of sulfonate in PbII complexes containing 5-sulfosalicylate is first reported in this presentation. The 4,4'-bipy ligands act as guest molecules in 1, dimeric linkers in 2, and monodentates in 3. The pi-pi stacking interactions can be observed in complexes 1-3, whereas there is no such interaction in complex 4. The coordination spheres of PbII ions in 1-4 are controlled by three factors: the activity of a lone pair of electrons, weak Pb-O interactions, and pi-pi stacking interactions. The PbII lone pair in 4 is inactive, whereas in 1-3, they are stereochemically active. The thermal stability and fluorescent property of complexes 1-4 are different from those of PbII complexes only containing chelating ligands, [Pb(Hssal)(2,2'-bipy)(DMF)]n (5), and [Pb(Hssal)(2,2'-bipy)(H2O)]n (6), and [Pb(Hssal)(phen)(DMF)]n (7).  相似文献   

12.
[Pb2(tfnb)4 (µ-CH3OH)] n (1) and [Pb2(dmp)2(tfnb)4] (2) (tfnb and dmp are the abbreviations for 4,4,4-trifluoro-1-naphthyl-1,3-butanedionate and 2,9-dimethyl-1,10-phenanthroline) have been synthesized and characterized by elemental analysis, IR, 1H NMR spectroscopy, and thermal analysis. The single-crystal structure of 1 shows that the complex forms two 1-D polymeric networks containing four types of Pb2+ with coordination numbers seven for Pb(1) and Pb(3), five for Pb(2), and six for Pb(4). The single-crystal structure of 2 shows that the complex forms a dinuclear complex with eight-coordinate Pb(II). The supramolecular features in this complex are guided by lone-pair activity and the control of weak directional intermolecular interactions and aromatic π–π stacking interactions.  相似文献   

13.
T Groh  K B?chmann 《Electrophoresis》1992,13(7):458-461
Organic and inorganic anions are analyzed using capillary zone electrophoresis with indirect fluorescence detection. As electrolyte 2,5-dihydroxybenzoic acid is used. By adding Pb2+ to the electrolyte the migration time of SO4(2-) can be selectively influenced and the separation of Cl- and SO4(2-) considerably improved. The fluorescence intensity of 2,5-dihydroxybenzoic acid increases in the presence of Pb2+ up to a concentration of 6*10(-4) mol/L Pb2+. Limits of detection in the lower mumol/L range were obtained.  相似文献   

14.
Hsu MH  Chen RT  Sheu WS  Shieh M 《Inorganic chemistry》2006,45(17):6740-6747
The novel hydroxo-bridged dimeric lead-chromium carbonyl complex [Et4N]2[{PbCr2(CO)10}2(mu-OH)2] ([Et4N]2[1]) was synthesized from the reaction of PbCl2 and Cr(CO)6 followed by metathesis with [Et4N]Br in a KOH/MeOH solution. The X-ray crystallographic structure shows that dianion 1 consists of two Pb{Cr(CO)5}2 units bridged by two hydroxo fragments in which the Pb atoms are further coordinated with two Cr(CO)5 groups, resulting in a distorted tetrahedral geometry. A CO2 molecule can insert itself into dianion 1 to form two new carbonate complexes, [Et(4)N]2[{PbCr2(CO)10}(CO3)] ([Et4N]2[2]) and [Et4N]2[{PbCr2(CO)10}2(CO3)] ([Et4N]2[3]), depending on the reaction conditions. In addition, complex 2 can be transformed into 3 in CH2Cl2 solution at an elevated temperature. While the carbonate group in dianion 2 is bonded to one Pb atom, which is coordinated with two Cr(CO)5 fragments, the carbonate group in 3 bridges the two Pb centers in a mu-1kappa2OO':2kappa2OO' fashion in which each Pb atom is further bonded to two Cr(CO)5 moieties. Complexes 2 and 3 can be converted back the hydroxo complex 1 under appropriate conditions. All three unprecedented lead-chromium compounds, 1-3, were fully characterized by spectroscopic methods and single-crystal X-ray diffraction analyses. The nature and formation of complexes 1-3 were also examined by molecular orbital calculations using the B3LYP method of the density functional theory.  相似文献   

15.
水热条件下,合成了2个含对位取代苯甲酸和2,2'-联吡啶配体的二价铅配合物[Pb(4-HOBA)2(2,2'-bipy)]·H2O(4-HOBAH=4-hydroxybenzoic acid,2,2'-bipy=2,2'-bipyridine)(1)及[Pb(4-MBA)2(2,2'-bipy)](4-MBAH=4-methylbenzoic acid)(2),并通过X-射线单晶结构分析、元素分析、红外光谱、紫外光谱、荧光光谱以及热重分析等手段对配合物进行了表征与研究.X-射线单晶结构测定表明,2种晶体的不对称单元中,六配位的二价铅离子分刖与来自1个2.2'-联吡啶的2个N原子和2个对位取代苯甲酸的4个O原子螫合.配合物1属单斜晶系,P21/c空间群,a=1.09483(4)nm,b=1.751 94(6)nm,c=1.20479(4)nm,β=100.334(2)°,V=2.27339(14)nm3,Z=4,final R1=0.0254,wR2=0.0674.广泛存在的分子间氢键和弱π…π堆积作用将配合物1组装成稳定的三维超分子结构.配合物2属三斜晶系,P-1空间群,a=0.955 10(11)nm,b=1.00805(12)am,c=1.324 83(15)nm,a=109.865(1)°,β=97.322(1)°,γ=90.643(1)°,V=1.1878(2)nm3,Z=2,final R1=0.0284,wR2=0.0508.配合物2经由一对Pb…O(0.3185(2)nm)弱相互作用被组装成二聚体结构.弱的Ph…O相互作用及与配合物1中类似的π…π堆积作用共同构筑了配合物2的三维结构.在配合物1和2中,铅的6s孤电子均显示了立体化学活性,使配位键分布于半球型区域.同时,两种配合物固体均具有光致发光特性,并在蓝光区有较强发射.  相似文献   

16.
The o-phenylenediamine-derived disulfonamide ligands 1 and 2 complex and efficiently extract Pb(II) from water into 1,2-dichloroethane via ion-exchange, in combination with 2,2'-bipyridine (97.5% and 95.0%, respectively, for 1:1 ligand-to-Pb ratios). The corresponding Pb(II)-sulfonamido binary complexes of ligands 1 and 2 (3 and 4, respectively), and ternary complexes with 2,2'-bipyridine (5 and 6, respectively), were isolated and characterized. (1)H NMR spectra of the organic phases after extraction show the formation of ternary Pb-sulfonamido-bipy complexes. X-ray characterization of 3, 4, and the ternary complex 5 consistently demonstrates four primary coordination sites and a stereochemically active lone pair on Pb. The X-ray structure of 3 shows a pseudo trigonal bipyramidal configuration on Pb, with the lone pair occupying one of the equatorial sites, and the formation of an unusual "hemidirected" coordination polymer via axial S=O-Pb coordination. The same axial S=O-Pb coordination pattern with two DMSO molecules is observed in the structure of 4.[2(CH(3))(2)SO)], thus rationalizing the high solubility of the binary complexes in strongly coordinating solvents. In contrast, the X-ray structure of the ternary complex 5 reveals a distorted four-coordinate configuration with only weak S=O-Pb coordination leading to dimer formation, thus explaining its higher solubility in weakly coordinating solvents. FT-IR spectroscopy confirms the X-ray data, since the ligand nu(S)(=)(O) stretching frequencies shift to lower values in the binary Pb(II)-sulfonamido complexes and are again altered upon formation of the ternary Pb(II)-sulfonamido-bipy complexes, as would be expected for 2,2'-bipy complexation and hindered S=O-Pb coordination.  相似文献   

17.
The stability constants of the 1:1 complexes formed between Pb2+ and the nucleosides (Ns), adenosine and guanosine, as well as between the nucleotides (NMP2-), AMP2-, IMP2-, and GMP2-, were determined by potentiometric pH titrations in aqueous solution (25 degrees C; I = 0.1 M, NaNO3). Based on previously established log KPb(R-PO3)Pb versus pKH(R-PO3)H straight-line plots (R-PO3(2-) = simple phosphate monoester or phosphonate ligands where R is a noninteracting site), it is shown that the Pb(IMP) and Pb(GMP) complexes are more stable than is expected on the basis of the basicity of the phosphate group of IMP2- and GMP2-. This means that macrochelates are formed, where the phosphate-coordinated Pb2+ also interacts with N7 of the nucleobase residue. In contrast, the stability of the Pb(AMP) complex is governed by the basicity of the AMP2- phosphate group. These results agree with the observations made for the Pb(Ns)2+ complexes: Pb(adenosine)2+ is very unstable in contrast to Pb(guanosine)2+, the stability of which is very similar to the one of Pb(cytidine)2+ studied previously. The stability constants of the Pb(Ns)2+ complexes also allowed an evaluation of the structure in solution of the monoprotonated Pb(H;NMP)+ complexes, the stabilities of which were also determined. We were able to show that the proton is located at the phosphate group and Pb2+ at the N7/(C6)O site of H(GMP)-; in the case of H(AMP)- Pb2+ is probably about equally distributed between the adenine residue and the monoprotonated phosphate group. On the basis of the stability constants of these complexes and their structures in solution, it is possible to provide a series which reflects the decreasing affinity for Pb2+ of nucleobase residues in single-stranded nucleic acids: guanine approximately equal to cytosine > (hypoxanthine) > adenine > uracil approximately equal to thymine. The Pb2+ affinity of the phosphodiester linkage, -PO3(-)-, is similar to the one of the adenine residue, but is expected to be more significant due to its larger abundance. The relevance of these results for lead-activated ribozymes is briefly discussed.  相似文献   

18.
Reaction of [Cp*Ir(dmpm)Cl]Cl with [Et3Si]B(ArF)4 (dmpm = bisdimethyl-phosphinomethane; ArF = C6F5) under hydrogen gas affords the dicationic complex [Cp*Ir(dmpm)H2]2+ (1), which is readily deprotonated by weak bases to give [Cp*Ir(dmpm)H]+. Complex 1 exists as a mixture of two isomers (97:3). On the basis of the magnitude of 2JH-P couplings and T1 measurements, a cis-dihydride or dihydrogen complex structure is suggested for the predominant isomer 1-cis (2JH-P = 6 Hz), with the minor isomer assigned a transoid structure 1-trans (2JH-P = 20 Hz). When the preparative reaction is carried out with HD gas, the resonance in the 1H NMR spectrum assigned to 1-cis-d1 exhibits1JH-D = 9.0 Hz. The observed values of 1JH-D vary significantly with temperature, increasing from 7.0 Hz at 223K to 9.0 Hz at 300 K. The observed chemical shift of 1-cis-d1 also varies significantly with temperature. These observations are interpreted in terms of a dynamic equilibrium between a cis-dihydride and a dihydrogen complex.  相似文献   

19.
The metal ion complexing properties of the ligand DPP (2,9-di-(pyrid-2-yl)-1,10-phenanthroline) were studied by crystallography, fluorimetry, and UV-visible spectroscopy. Because DPP forms five-membered chelate rings, it will favor complexation with metal ions of an ionic radius close to 1.0 A. Metal ion complexation and accompanying selectivity of DPP is enhanced by the rigidity of the aromatic backbone of the ligand. Cd2+, with an ionic radius of 0.96 A, exhibits a strong CHEF (chelation enhanced fluorescence) effect with 10(-8) M DPP, and Cd2+ concentrations down to 10(-9) M can be detected. Other metal ions that cause a significant CHEF effect with DPP are Ca2+ (10(-3) M) and Na+ (1.0 M), whereas metal ions such as Zn2+, Pb2+, and Hg2+ cause no CHEF effect with DPP. The lack of a CHEF effect for Zn2+ relates to the inability of this small ion to contact all four donor atoms of DPP. The structures of [Cd(DPP)2](ClO4)2 (1), [Pb(DPP)(ClO4)2H2O] (2), and [Hg(DPP)(ClO4)2] (3) are reported. The Cd(II) in 1 is 8-coordinate with the Cd-N bonds to the outer pyridyl groups stretched by steric clashes between the o-hydrogens on these outer pyridyl groups and the central aromatic ring of the second DPP ligand. The 8-coordinate Pb(II) in 2 has two short Pb-N bonds to the two central nitrogens of DPP, with longer bonds to the outer N-donors. The coordination sphere around the Pb(II) is completed by a coordinated water molecule, and two coordinated ClO4(-) ions, with long Pb-O bonds to ClO4(-) oxygens, typical of a sterically active lone pair on Pb(II). The Hg(II) in 3 shows an 8-coordinate structure with the Hg(II) forming short Hg-N bonds to the outer pyridyl groups of DPP, whereas the other Hg-N and Hg-O bonds are rather long. The structures are discussed in terms of the fit of large metal ions to DPP with minimal steric strain. The UV-visible studies of the equilibria involving DPP and metal ions gave formation constants that show that DPP has a higher affinity for metal ions with an ionic radius close to 1.0 A, particularly Cd(II), Gd(III), and Bi(III), and low affinity for small metal ions such as Ni(II) and Zn(II). The complexes of several metal ions, such as Cd(II), Gd(III), and Pb(II), showed an equilibrium involving deprotonation of the complex at remarkably low pH values, which was attributed to deprotonation of coordinated water molecules according to: [M(DPP)(H2O)]n+ <==> [M(DPP)(OH)](n-1)+ + H+. The tendency to deprotonation of these DPP complexes at low pH is discussed in terms of the large hydrophobic surface of the coordinated DPP ligand destabilizing the hydration of coordinated water molecules and the build-up of charge on the metal ion in its DPP complex because of the inability of the coordinated DPP ligand to hydrogen bond with the solvent.  相似文献   

20.
王伟  寇元  田中庸裕 《化学学报》2003,61(2):295-297
在超稀浓度条件下ω[Pb(OAc)2]或ω[Pb(OAc)4]<0.2,我们使用EXAFS方法对 溶解在离子液体中的铅化合物的局域结构进行了研究。从结果中可以看出,Pb (OAc)4的铅中心在(0.200±0.003)nm处有8个氧近邻,而Pb(OAc)2的铅中心在( 0.225±0.003)nm处有6个氧近邻。但是溶解在离子液体中的铅化合物均具有较低的 配位数(一般小于4),这显示出离子液体具有较低的配位性能。另一方面,铅中 心与最近邻氧的原子间距有较大的变化。例如:溶解于[bmim]PF6的Pb^4+的Pb-O间 距为(0.186±0.005)nm,配位数仅为2.4,从Fourier变换谱上可清楚地看出有第 二配位壳层的贡献;溶解于[bmim]PF6的Pb^2+的Pb-O间距也明显地小于纯样品中 Pb-O的间距。相比之下,溶解于[ bmim]BF4的铅离子显示出的Pb-O间距与纯样品几 乎没有差别,对于Pb^2+为0.223nm,而对于Pb^4+为0.210nm。  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